You are on page 1of 24

Geoscience Frontiers 12 (2021) 549–572

H O S T E D BY Contents lists available at ScienceDirect

Geoscience Frontiers
journal homepage: www.elsevier.com/locate/gsf

Research Paper

Depositional age and tectonic environment of the Gouap banded iron


formations from the Nyong group, SW Cameroon: Insights from isotopic,
geochemical and geochronological studies of drillcore samples
Landry Soh Tamehe a, b, Chongtao Wei a, *, Sylvestre Ganno b, Carlos Alberto Rosiere c,
Jean Paul Nzenti b, Chesther Gatse Ebotehouna d, Guanwen Lu a
a
Key Laboratory of CBM Resources and Reservoir Formation Process, Chinese Ministry of Education, China University of Mining and Technology, Xuzhou, Jiangsu,
221116, China
b
Department of Earth Sciences, University of Yaounde I, P.O. Box 812, Yaounde, Cameroon
c
Department of Geology, Universidade Federal de Minas Gerais, 31270-901, Belo Horizonte, Minas Gerais, Brazil
d
School of Civil and Environmental Engineering, University of Science and Technology Beijing, Beijing 100083, China

A R T I C L E I N F O A B S T R A C T

Handling Editor: Christopher Spencer The discovery of the Gouap banded iron formations (BIFs)-hosted iron mineralization in the northwestern of the
Nyong Group (Ntem Complex) in southwestern Cameroon provides unique insights into the geology of this region.
Keywords: In this contribution, we firstly report detailed study of geochemistry, isotopic and geochronology of well pre-
Gouap served samples of the Gouap BIFs collected from diamond drillcores. The Gouap BIFs consist mainly of amphibole
Banded iron formations
BIFs and amphibole–pyrite BIFs characterized by dominant Fe2O3 þ SiO2 contents and variable contents of CaO,
Isotope geochemistry
MgO and SO3, consistent with the presence of amphibole, chlorite, epidote and pyrite, formed during amphibolite
Zircon geochronology
Ryacian period facies metamorphism and overprinted hydrothermal event. The amphibole–pyrite BIFs are typically enriched in
Brasiliano–Congo orogeny trace and rare earth elements (REE) compared to the amphibole BIFs, suggesting the influence of detrital materials
as well as secondary hydrothermal alteration. The Post Archean Australian Shale (PAAS)-normalized REE–Y
profiles of the Gouap BIFs display positive La, Eu anomalies, weak negative Ce anomalies, indicating a mixture of
low-temperature hydrothermal fluids and relatively oxic conditions probably under relative shallow seawater.
We present here the first isotopic data of BIFs within the Ntem Complex. The δ30SiNBS28 values of the quartz
from the Gouap BIFs vary from 1.5‰ to 0.3‰ and from 0.8‰ to 0.9‰ for the amphibole BIFs and
amphibole–pyrite BIFs, respectively. The quartz has δ18OV-SMOW values of 6.8‰–9.5‰ (amphibole BIFs) and
9.2‰–10.6‰ (amphibole–pyrite BIFs). The magnetite from the Gouap BIFs shows δ18O values ranging from
3.5‰ to 1.8‰ and from 3‰ to 1.7‰ for the amphibole BIFs and amphibole–pyrite BIFs, respectively.
Moreover, the pyrite grains in the amphibole–pyrite BIFs display δ34S values of 1.1‰–1.8‰. All isotopic data of
the Gouap BIFs confirm that they might have precipitated from low-temperature hydrothermal fluids with detrital
input distant from the volcanic activity. According to their geochemical and isotopic characteristics, we propose
that the Gouap BIFs belong to the Superior type.
In situ U–Pb zircon dating of BIFs was conducted to assess the BIF depositional age based on strong evidence of
zircon in thin section. The Gouap BIFs were probably deposited at 2422  50 Ma in a region where sediments
extended from continental shelf to deep-water environments along craton margins like the Cau^e Formation of the
Minas Supergroup, Brazil. The studied BIFs have experienced regional hydrothermal activity and metamorphism
at 2089  8.3 Ma during the Eburnean–Transamazonian orogeny. These findings suggest a physical continuity
between the protocratonic masses of both S~ao Francisco and Congo continents in the Rhyacian Period.

* Corresponding author.
E-mail addresses: weighct@163.com (C. Wei), sganno2000@yahoo.fr (S. Ganno).
Peer-review under responsibility of China University of Geosciences (Beijing).

https://doi.org/10.1016/j.gsf.2020.07.009
Received 27 August 2019; Received in revised form 22 April 2020; Accepted 2 July 2020
Available online 25 August 2020
1674-9871/© 2020 China University of Geosciences (Beijing) and Peking University. Production and hosting by Elsevier B.V. This is an open access article under the
CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

1. Introduction of oxide-facies, characterized by rhythmic and rarely irregular alterna-


tion of granular or recrystallized quartz-rich and iron-rich layers. These
The banded iron formations (BIFs) are one of the key rocks to get BIFs exhibit strong deformational features and they are spatially associ-
insightful information of the Earth’s history. In fact, the BIFs occur in ated with different country rocks (e.g., gneiss, quartzite, schist,
Precambrian terrains throughout the world and therefore they play a amphibolite, syenite, epidotite), which both have undergone a greens-
significant role for understanding the composition of the Earth’s hydro- chist to granulite facies metamorphic grade (Ganno et al., 2015, 2016,
sphere and atmosphere as well as the evolution of the Earth’s crust 2017, 2018; Teutsong et al., 2017; Ndime et al., 2018, 2019; Soh Tamehe
(Trendall, 2002; Huston and Logan, 2004; Bekker et al., 2010; Kato et al., et al., 2018, 2019). Previous studies have been attempted to characterize
2006; Soh Tamehe, 2020). These explain the several researches that have the features and genesis of the metamorphic BIFs of the Ntem Complex
been conducted worldwide on the BIFs during the past decades. Despite (e.g., Ganno et al., 2017; Teutsong et al., 2017; Soh Tamehe et al., 2018;
the abundant literature on BIFs dealing with their mineralogy, petrology, Ndime et al., 2018, and references therein). Nevertheless their deposi-
geochemistry and geochronology, the BIFs remain enigmatic rocks while tional settings remain to be precisely constrained. Earlier authors have
constituting an important source of iron ore for the steel industry classified these rocks as being deposited in a volcanic environment
nowadays. (Algoma-type; Teutsong et al., 2017; Nzepang Tankwa et al., 2020) or on
The Precambrian Ntem Complex corresponds to the northern edge a passive margin (Superior-type; Ganno et al., 2015, 2017; Soh Tamehe
extension of the Congo craton in southern Cameroon (Maurizot et al., et al., 2018) according to their petrographical and geochemical
1986). The Ntem Complex has usually been subdivided into the Nyong characteristics.
and Ntem Groups, where iron deposits are hosted by metamorphic BIFs Previous studies on metamorphic BIFs in the Nyong Group were
(e.g., Suh et al., 2009; Chombong and Suh, 2013; Ilouga et al., 2013; carried out mostly on surface weathered outcrops (Ganno et al., 2016,
Ganno et al., 2015, 2016). The BIFs of the Ntem Complex mainly consist 2017; Soh Tamehe et al., 2018). Besides the reliable outcomes achieved

Fig. 1. Geological map of the SW Cameroon showing the location of the studied area (blue square) and major BIF-hosted iron deposits (red circle) (adapted from
Maurizot et al., 1986). Inset illustrating the position of the SW Cameroon (red rectangle) relative to the Congo craton in Africa.

550
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

for their studies, they cannot record the expected information for un- This paper presents the first detailed geochemical and geochrono-
derstanding their origin, whereas drillcore allow us to collect well pre- logical studies of well preserved samples of the Gouap BIFs collected from
served samples to better constrain the genesis as well as the depositional diamond drillcores in the Gouap iron deposit, which represents a pro-
environment of the BIFs. Drilling exploration programs have been con- spective area for iron ore in southwestern Cameroon. Furthermore, this
ducted by G-Stones Resources Ltd. at the Gouap prospect in the northwest study provides the first isotopic data of quartz, magnetite and pyrite of
of the Nyong Group (Fig. 1), providing drillcore samples that displays the the BIFs within the Ntem Complex. Direct dating of zircon extracted from
interlayering of the iron formations with gneisses, schists, quartzites and the Gouap BIFs was achieved using laser ablation inductive coupled
dolerite sills (Soh Tamehe et al., 2019). Limited geological investigations plasma mass spectrometry (LA-ICP-MS). These new results aim to
have so far been accomplished on BIF core samples from the Nyong constrain the genesis and depositional age of the Gouap BIFs as well as
Group (Soh Tamehe et al., 2019; Nzepang Tankwa et al., 2020) compared the tectonic environment from which the BIFs were precipitated.
to those from the Ntem Group (Teutsong et al., 2017; Ndime et al., 2018, Thereby, this study is a substantial contribution towards the improved
2019). understanding of the BIF genesis of the Nyong Group.

Fig. 2. (a) Detailed geological map of the Nyong Group showing the location of the Gouap prospect (black square) and BIF-hosted iron deposits (red circle) (modified
after Lerouge et al., 2006). (b) Sketch map of the Congo and S~ao Francisco cratons (greyed area) with limits of NE Brazil and Central West Africa illustrating location of
the studied area (red star).

551
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

2. Geological background concentrated on the hinge zones of the F2 folds, while the hema-
tite–martite and magnetite–hematite bodies developed on highly frac-
2.1. The Nyong Group tures zones.

The Gouap iron deposit belongs to the Nyong Group (Fig. 2), which 3. Samples description and analytical methods
consists mainly of metasedimentary and metavolcanic rocks as well as
syn- to late-tectonic intrusive rocks. It includes various migmatitic rocks, 3.1. Samples description
gneisses, mafic granulites, charnockites, pyroxenites, amphibolites,
schists, quartzites and BIFs (e.g., Toteu et al., 1994; Lerouge et al., 2006; For this study, we focused on the two major BIF lithofacies (amphi-
Ndema et al., 2014; Ganno et al., 2017). Intrusions of metadiorites, bole BIFs and amphibole–pyrite BIFs) of the Gouap deposit. The BIF
granodiorites, and syenites cut rocks of the Nyong Group, along with samples were collected from different depths of three drillholes: DD036
post-tectonic dolerites (Toteu et al., 1994; Lerouge et al., 2006). More (length ¼ 151.50 m, azimut ¼ 115 and dip ¼ 55 ), DD042 (length ¼
recently, well-preserved to variably retrogressed eclogite facies meta- 200.90 m, azimut ¼ 115 and dip ¼ 60 ), and DD043 (length ¼ 183.00
morphic rocks (eclogites and serpentinitized peridotites) have been re- m, azimut ¼ 115 and dip ¼ 55 ). These drillholes are located in the
ported in the Nyong Group (Loose and Schenk, 2018; Bouyo southern part of the Gouap iron deposit (Fig. 3), and they are represen-
Houketchang et al., 2019; Nga Essomba et al., 2020). tative drillholes in terms of the geology of the deposit (Soh Tamehe et al.,
The metasedimentary and metaplutonic rocks of the Nyong Group 2019). The studied BIF samples are all fresh and free from weathering.
have undergone a granulite facies metamorphism associated with the The Gouap BIF samples collected are described below.
intrusion of charnockitic rocks (Toteu et al., 1994; Lerouge et al., 2006). In hand specimen, samples of amphibole BIF from the Gouap deposit
Toteu et al. (1994) reported U–Pb zircon metamorphic ages (~2050 Ma) are slightly dark gray to dark green in color (Fig. 5a and b). The rocks are
from gneissic rocks at Kribi and Akom II (see Fig. 2). Subsequently, the fine-grained and display a well-developed banding consisting of inter-
high-grade metamorphic event (2100–2000 Ma) associated with tecto- calated laminations of thinner silica-rich (1–3 mm thick) and oxide-rich
nomagmatic activities (2055  5 Ma) were established using the Sensi- (0.5–2 mm thick) layers (Fig. 5a and b). On a microscopic scale, the
tive High Resolution Ion-Microprobe (SHRIMP) U–Pb analyses on zircon amphibole BIF exhibits granoblastic texture with abundant magnetite,
grains from metasedimentary and metaplutonic rocks collected at Lolo- quartz and amphibole grains, and minor epidote and zircon (Fig. 5c–f).
dorf (Lerouge et al., 2006). The emplacement of these plutonic rocks The silica-rich layers are dominated by amphibole and quartz, whereas
within the Nyong Group is associated with Eburnean high-grade meta- the oxide-rich layers are made up of magnetite (Fig. 5c–f). Magnetite
morphism which continued probably up to 1985  8 Ma (Lerouge et al., forms coarse (2.3 mm  0.7 mm) euhedral to subhedral crystals with
2006). In addition, these authors (Toteu et al., 1994; Lerouge et al., 2006) preferred orientation and is typically surrounded by quartz and amphi-
identified other non-Archean events recorded by Eburnean zircon growth bole (Fig. 5cd). Magnetite often contains quartz inclusions. Quartz
and titanite ages in metagranitoids. Thus, the Nyong Group recorded appears as fine to medium-grained subhedral to euhedral crystals (0.7
Paleoproterozoic and Pan-African reworking. More recently, Chombong mm  0.3 mm), and generally contains inclusions of magnetite and
et al. (2017) obtained Neoarchean age (2699  7 Ma) and Neo- zircon (Fig. 5c–f). Amphibole mostly occurs as fine subhedral to euhedral
proterozoic age (600–500 Ma) from zircon of magnetite gneisses. The crystals (0.1–0.5 mm), and intergrowths with quartz. Zircon is scarce and
former age represents the maximum age of formation of the magnetite occurs as very fine-grained (<0.1 mm) euhedral crystals in samples such
gneisses, whereas the later age is interpreted as the Pan-African distur- as GB11–2 and GB9 (Fig. 5c–f).
bance recorded by these gneisses (Chombong et al., 2017). At the Gouap deposit, samples of the amphibole–pyrite BIF display
Additionally, Loose and Schenk (2018) demonstrated that the Nyong dark gray to dark green color and banding structure in hand specimen
Group underwent an eclogite facies metamorphism at 2093  45 Ma (Fig. 6a and b). The rocks are coarse-grained and their primary sedi-
based on SHRIMP U–Pb dating performed on zircon grains from eclogites mentary micro-banding is obliterated and tectonically transposed with
collected near the Bipindi area. This age is comparable to that of meta- the development of metamorphic foliation characterized by alternating
morphic events reported earlier by Toteu et al. (1994) and Lerouge et al. light silica-rich (1–2.5 mm thick) and dark oxide-rich (0.5–1.5 mm thick)
(2006). This implies that the Nyong Group experienced similar meta- bands (Fig. 6a and b). In thin section, the amphibole–pyrite BIF is char-
morphism during the Paleoproterozoic between 2100 and 2000 Ma acterized by heterogranular granoblastic texture made mainly of quartz,
(Bouyo Houketchang et al., 2019). magnetite, amphibole, pyrite, and minor clinopyroxene, epidote, chlorite
The Nyong Group displays flat-lying foliations with N–S to NNE–SSW and zircon. Quartz and amphibole composed the silica-rich laminae,
axial fold lineations and E–W to NW–SE stretching lineations that indi- while magnetite and pyrite constitute the oxide-rich laminae (Fig. 6c–f).
cate an eastward-directed movement (Maurizot et al., 1986). Quartz occurs as fine- to medium-grained anhedral crystals (0.1–0.9
mm), and commonly contains inclusions of magnetite and zircon
2.2. The Gouap iron deposit (Fig. 6c–f). Amphibole forms subhedral to anhedral laths crystals
(0.3–0.5 mm) which is often altered to chlorite (Fig. 6c–f). Magnetite
Detailed stratigraphical, petrographical and structural information of grains (0.3–1.4 mm) appears as elongated subhedral to euhedral crystals
the Gouap iron deposit are given by Soh Tamehe et al. (2019). The and contains quartz inclusions (Fig. 6c–f). Pyrite generally occurs as fine-
geological sketch map of the deposit is presented in Fig. 3. The Gouap to medium-grained (0.06–0.3 mm) euhedral to subhedral crystals, and is
iron deposit is comprised of amphibole BIFs and amphibole–pyrite BIFs, disseminated in the rock matrix or is aggregated with magnetite
well interlayered with amphibolite facies gneisses (garnet, amphib- (Fig. 6c–f). Secondary chlorite and epidote flakes (0.1–0.3 mm) occur as
ole–magnetite and amphibole–garnet–pyrite gneiss), staurolite micas- an alteration product of biotite and amphibole, respectively (Fig. 6c and
chists and quartzites displaying sharp to gradational contact. Younger d). Clinopyroxene appears as fine-grained (0.6–0.8 mm) subhedral to
dolerite sills crosscut the gneisses and BIFs. The deposit has experienced anhedral crystals, and is often associated with magnetite (Fig. 6c–f).
three principal deformations events that led to the development of foli- Zircon grains (<0.1 mm) commonly appears as euhedral to subhedral
ation, folding and faulting on a regional scale. crystals in quartz grain (Fig. 6c–f).
The Gouap iron deposit hosts both soft and hard iron ore types. The
soft iron ore type, which dominantly occurs in the western and northern 3.2. Major and trace elements geochemistry
areas of the Gouap deposit, is comprised of hematite–martite and mag-
netite–hematite, whereas the hard type is magnetitic and mainly found in Sixteen BIF samples (eight amphibole BIFs and eight amphib-
the southern part of the deposit. The magnetite ore bodies are mostly ole–pyrite BIFs) were collected from different depths of drillholes

552
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 3. Geological map of the Gouap iron deposit (adapted after Soh Tamehe et al., 2019).

553
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 4. Stratigraphic columns showing the location of the BIF samples collected from the drillholes of the Gouap iron deposit (modified after Soh Tamehe et al., 2019).

DD036, DD042 and DD043 (see Fig. 3) and used for geochemical ana- platinum mould. The resultant disk is in turn analyzed by X–ray fluo-
lyses. The locations of the BIF samples are shown in Fig. 4a–c. Major and rescence (XRF) spectrometry. The XRF analyses are determined in
trace elements analyses were performed at the Australian Laboratory conjunction with a loss-on-ignition at 1000  C. The resulting data from
Services (ALS) of Guangzhou, China. For major elements, a prepared both determinations are combined to produce a “total”.
sample is fused with lithium metaborate-lithium tetraborate flux which For trace elements, a prepared sample is added to lithium metaborate-
also includes an oxidizing agent (Lithium Nitrate), and then poured into a lithium tetraborate flux, mixed well and fused in a furnace at 1025  C.

554
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 5. Drillcore samples and photomicrographs of the amphibole BIF from the Gouap iron deposit. (a, b) Typical hand specimen consisting of oxide-rich layer
alternating with silica-rich layer. (c, d) Granoblastic texture showing a crystallographic preferred orientation of magnetite and quartz coexisting with amphibole and
zircon inclusion (plane polarized light). (e, f) Magnetite megacryts surrounded by quartz megacrysts (reflected light). Abbreviations: Amp, amphibole; Mag, magnetite;
Qtz, quartz; Zr, zircon.

The resulting melt is then cooled and dissolved in an acid mixture con- All the isotopic analyses were performed at the Analytical Laboratory
taining nitric, hydrochloric and hydrofluoric acids. This solution is then Beijing Research Institute of Uranium Geology (ALBRIUG), Beijing,
analyzed by ICP-MS. Analytical uncertainties vary from 0.1% to 0.04% China.
for major elements; and 0.1%–0.5% for trace elements. Various standards For oxygen isotope analyses, quartz and magnetite grains were
were used and data quality assurance was verified by running these crushed into 200 mesh and reacted with pure BrF5 during 14 h at
standards between samples as unknowns. The detection limits were 0.01 500–680  C under vacuum pressure of 103 Pa. The reaction generated
ppm for the REE and Y. O2 and impurity components (SiF4 and BrF3), which was separated by
freezing method. Then pure O2 was reacted with a carbon rod at 700  C,
3.3. Isotopes geochemistry producing CO2 which was collected by freezing method and analyzed by
MAT–253 mass spectrometer for oxygen isotope compositions. The result
Quartz and magnetite grains were separated from nine BIF samples. of the measurement is expressed as δ18OV-SMOW in parts per mil (‰)
One amphibole BIF (GB9) sample was collected from drillhole DD036, based on the Vienna Standard Mean Ocean Water (V–SMOW). The ac-
whereas three amphibole–pyrite BIF (GB10–1, GB10–2 and GB10-3) and curacy of the analyses was 0.2‰.
five amphibole BIF (GB11-1 to GB11-5) samples were collected from Silicon isotope analyses of quartz were preceded after quartz grains
drillhole DD043 (see location in Fig. 4b and c). Quartz grains were were crushed into 200 mesh and reacted with pure BrF5 at 500–680  C
separated prior to silicon and oxygen isotope analyses, while magnetite for 14 h under 103 Pa vacuum pressure. The reaction released SiF4
grains were separated only prior to oxygen isotope analyses. Prior to which was separated, collected, and purified using dry ice, liquid nitro-
sulfur isotope analyses, the pyrite grains were extracted from ten BIF gen and alcohol, respectively. The Si isotope composition was measured
samples (GB10-1 to GB10-10, see location in Fig. 4c) collected from by MAT–253 mass spectrometer. The measurement result is expressed as
drillhole DD043. Quartz, magnetite and pyrite grains were handpicked δ30SiNBS-28 in parts per mil (‰) relative to the NBS–28 as standard. The
from the 60–80 mesh mineral crush under a binocular (purity > 99%). analytical uncertainties were better than 0.1‰.

555
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 6. Drillcore samples and photomicrographs of the amphibolepyrite BIF from the Gouap iron deposit. (a, b) Typical hand specimen showing pyrite grains within
oxide-rich layer. (c, d) Granoblastic texture displaying magnetite coexisting with pyrite in oxide-rich layer (plane polarized light). (e, f) Intergrown Py þ Mag þ Qtz
with minor Amp þ Cpx and Zr inclusion (reflected light). Abbreviations: Amp, amphibole; Bt, biotite; Cpx, clinopyroxene; Mag, magnetite; Qtz, quartz; Py, pyrite;
Zr, zircon.

For sulfur isotope analyses, pyrite grains were crushed into 200 mesh structures and select appropriate spots prior to U–Pb dating. Back-
and mixed uniformly with cuprous oxide. The reaction was heated at 980 scattered electron (BSE) images of zircon grains were further taken to

C under 2  102 Pa vacuum pressure. The SO2 was generated and better interpret their internal fabrics. CL and BSE images were collected
collected by freezing method, then analyzed by a MAT–251 mass spec- using a JSM 6510 scanning electron microscope (SEM; JEOL) attached to
trometer for sulfur isotopes. The measurement result is reported as δ34SV- a Gatan CL detector with a voltage of 10 kV.
CDT in parts per mil (‰) based on the Vienna Ca~ non Diablo Troilite Zircon U–Pb analyses were performed using an ESI NWR 193 nm laser
(V–CDT) standard. The accuracy of the analyses was better than 0.2‰. ablation system and an AnalytikJena PQMS Elite ICP-MS instrument. The
zircon grains were analyzed using a 32 μm spot diameter. All analyses
3.4. Zircon U–Pb dating were conducted with a 10 Hz repetition rate and energy of 2.5 J cm2.
For every set of 5–10 analyses, five reference standards were used: (1)
Two representative BIF samples (GB10-1 and GBW3-1) collected from NIST610 glass (http://georem.mpch-mainz.gwdg.de/), (2) 91500 (Wie-
drillholes DD042 and DD043 were used for dating (see location in denbeck et al., 1995), (3) GJ1 (Jackson et al., 2004), (4) Plesovice (Slama
Fig. 4a–c). Zircon grains were extracted from each sample using con- et al., 2008), and (5) Qinhu (Li et al., 2013). This was done to correct for
ventional gravimetric and magnetic techniques, following by hand- laser-induced elemental fractionation, instrumental mass fractionation
picking under a binocular microscope at Langfang Rock Detection and drift, to calibrate the measured isotopic ratios, and to validate the
Technology Services Ltd (Hebei, China). Then all subsequent zircon measurement procedure (Quinn et al., 2016). The ICPMSDataCal 8.4
treatment was conducted at the Beijing Kehui Testing International Co. software was used for fractionation correction and calculation of the
Ltd, China. Representative zircon grains were mounted in epoxy resin dating results (Liu et al., 2010). The 207Pb/206Pb, 206Pb/238U, 207U/235U
and polished to mid-section to expose their cores for analyses. Cath- and 208Pb/232Th isotopic ratios were corrected using zircon 91500 as an
odoluminescence (CL) images, transmitted and reflected light micro- external standard for both instrumental bias and elemental and isotopic
photographs of all zircon grains were taken to examine their internal fractionation. Common lead was corrected using the method of Anderson

556
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

(2002). Uncertainties in the age analysis are quoted at 1σ, whereas errors

105.05
in the weighted mean ages are calculated at 2σ level (95% confidence).

53.65

32.31

85.96
Mean

0.23

3.93
0.23
3.15
2.43
0.20
0.39
0.17
8.42
2.59

4.17
Concordia diagrams, weighted average age calculation and probability
density plotting were processed using Isoplot/Ex 3.0 (Ludwig, 2003).

GB10-2
171.35
DD043
A detailed description for LA-ICP-MS instrument is presented in Ap-

47.21

36.80

99.74
84.01
0.25

5.86
0.07
3.01
3.51
0.21
0.95
0.37
1.50
0.42

6.11
pendix S1, whereas full analytical data for zircon grains (analyzed spots
and standards) is given in Appendix S2. This follows the recommenda-
tions outlined by Kosler et al. (2013) and Horstwood et al. (2016). To

GB10-1
171.30
DD043

44.79

39.94

99.55
84.73
0.29

5.16
0.07
3.27
3.07
0.24
0.96
0.44
1.32
0.95

5.45
calculate the most geologically appropriate age from our U–Pb zircon
geochronological data, we also follow the recommendations by Spencer
et al. (2016). However, the maximum depositional age (MDA) of the BIF

160.00
DD042

55.40

36.66

99.84
92.06
GB6*
samples was calculated as the weighted average of the youngest detrital

0.11

2.14
0.07
2.29
1.68
0.10
0.58
0.31
0.50
0.03

2.25
zircon population (minimum of n ¼ 3) that overlap within uncertainty at
the 2σ level (Dickinson and Gehrels, 2009). The MDA for sample GB10-1

GBW3-5
was further calculated after excluding data of overdispersed grain spots.

184.50

111.45
DD042

62.08

21.98

16.70

84.06
0.16

4.76
0.32
2.68
2.01
0.61
0.11
0.04

4.71

4.92
4. Results

GBW3-4
181.10
DD042

<0.01
67.03

24.25

99.32
91.28
4.1. Major elements

0.01

0.42
0.19
3.68
3.29
0.08

0.04
0.33
0.48

0.43
The major elements analyses of the Gouap BIFs are presented in

GBW3-3
Table 1. The amphibole BIFs are characterized by high Fe2O3

177.00

102.59
DD042

60.65

28.83

89.48
0.01

0.81
0.29
3.82
3.25
0.09
0.02
0.07
4.75
1.23

0.82
(44.57–60.83 wt.%, expressed as Fe total) and SiO2 (36.22–50.12 wt.%)
contents, which represent the major components (>91 wt.%) of the bulk
rock (Table 1). The Al2O3 and TiO2 have low concentrations ranging from

GBW3-2*
Amphibole–pyrite BIF
0.15 to 3.23 wt.% and <0.01 to 0.10 wt.%, respectively. Other oxides

151.54

107.74
DD042

58.22

26.27

10.75

84.49
0.43

6.71
0.44
3.28
1.47
0.12
0.04
0.01

2.81

7.14
such as MgO (2.19–3.85 wt.%) and CaO (0.71–2.28 wt.%) display sig-
nificant contents in the studied BIFs, which is consistent with the pres-
ence of amphibole. Very low concentrations of MnO (0.04–0.18 wt.%), GBW3-1*
144.55

120.18
DD042
Na2O (0.02–0.34 wt.%), P2O5 (0.17–0.26 wt.%), and SO3 (0.02–0.24

33.84

43.76

10.11

77.60
0.59

5.61
0.36
3.15
1.12
0.13
0.07
0.05
31.5

6.20
wt.%) characterize these amphibole BIFs. Most BIF samples except for
GB4-3 and GB11-2 show K2O contents below the detection limit (<0.001
wt.%). Loss on ignition (LOI) is very low for all the amphibole BIF

100.57
41.56

53.53

95.09
Mean

0.04

0.85
0.08
3.01
1.17
0.09
0.07
0.20
0.07
1.10

0.89
samples and ranges from 1.51 to 0.80 wt.%.
The amphibole–pyrite BIFs consist principally of high SiO2 content
(33.84–67.03 wt.%) and relatively low content of Fe2O3 (21.98–43.76
129.50

100.36
DD036

<0.01
50.12

44.57

94.69
0.01

0.20
0.18
2.71
2.28
0.04

0.19
0.06
0.80

0.21
GB9

wt.%, as Fe total), when compared to amphibole–BIFs. The TiO2 content


is generally low, ranging from 0.01 to 0.59 wt.%, while the Al2O3 content
Major elements compositions (wt.%) of the Gouap BIFs. *Data from Soh Tamehe et al. (2019).

(0.42–6.71 wt.%) is noticeably high in these BIFs. Concentrations of CaO


GB11-2
104.00

101.03
DD043

43.32

48.64

91.96
and MgO are also elevated within the amphibole–pyrite BIFs, ranging
0.10

3.23
0.13
3.85
1.14
0.34
0.08
0.18
0.02
0.85

3.33
from 1.12 to 3.51 wt.% and 2.29 to 3.82 wt.%, respectively. This cor-
roborates the occurrence of Mg-hornblende, tremolite and cumming-
GB11-1
103.00

100.29
DD043

tonite in these BIFs (Soh Tamehe et al., 2019). The studied BIFs also have
<0.01
37.81

58.45

96.26
0.03

0.71
0.05
2.19
0.71
0.05

0.26
0.03
1.13

0.74
significant SO3 contents (0.33–16.70 wt.%), which confirm the presence
of pyrite. All other oxides such as MnO (0.07–0.44 wt.%), Na2O
(0.08–0.24 wt.%), K2O (<0.01–0.96 wt.%), and P2O5 (0.01–0.44 wt.%)
154.05

100.96
DD042
GB4-5

<0.01

<0.01

<0.01
36.22

60.83

97.05
0.17
0.04
2.73
0.72
0.03

0.22

1.51

0.17

have very low contents. The amphibole–pyrite BIFs display low to high
LOI varying from 0.03 to 10.11 wt.%. The samples with high LOI also
show high SO3 contents, suggesting highly hydrothermal alteration of
147.00

100.41
DD042
GB4-4

<0.01

<0.01

<0.01
41.47

54.25

95.72
0.15
0.05
3.50
0.77
0.03

0.19

1.02

0.15

these rocks.

4.2. Trace and rare earth elements


GB4-3*
136.50

100.16
DD042

40.24

52.32

92.56
0.06

1.78
0.08
3.71
1.60
0.10
0.05
0.20
0.02
0.90

1.84

The trace and rare earth elements analyses of the Gouap BIFs are given
Note: Fe2OT3, total Fe expressed as Fe2O3.

in Table 2. The amphibole BIFs particularly have very low (<5 ppm) to
GB4-2*
133.10

100.55
DD042

moderate (<40 ppm) concentrations of trace elements when compared to


<0.01

<0.01
42.28

53.97

96.25
0.22
0.04
2.51
1.10
0.02

0.17
0.24
1.16

0.22

the average composition of the upper continental crust from Rudnick and
Amphibole BIF

Gao (2003) (Fig. 7a). Except Zn (15–67 ppm) and Cr (10–30 ppm) con-
tents, the amphibole BIFs are characterized by low concentrations (<20
GB4-1*
132.05

100.78
DD042

<0.01
41.04

55.18

96.22
0.01

0.30
0.05
2.88
1.01
0.08

0.20
0.03
1.44

0.31

ppm) of transition trace metals. In terms of large ion lithophile elements


(LILE), these BIFs have very low contents (<5 ppm) except for Sr and Ba
contents varying between 9.5–39.4 ppm and 1.0–24.8 ppm, respectively.
Fe2OT3 þ SiO2
Al2O3 þ TiO2

The high field strength elements (HFSE) display also very low concentra-
Rock code
Depth (m)
Drill hole
Lithology

tions (<5 ppm) in the amphibole BIFs, except for Zr which reaches values
Fe2O3T
Table 1

Al2O3

Na2O

Total
MnO

P2O5
MgO
TiO2
SiO2

CaO

K2O

SO3

up to 27 ppm. The contents of other trace elements (e.g., Li, Be and Ga) are
LOI

trivial in all these BIF samples (Table 2, Fig. 7a). Most amphibole BIF

557
L. Soh Tamehe et al.
Table 2
Trace and rare earth elements (ppm) of the Gouap BIFs.
Lithology Amphibole BIF Amphibole–pyrite BIF

Rock code GB4-1 GB4-2 GB4-3 GB4-4 GB4-5 GB11-1 GB11-2 GB9 Mean GBW3-1 GBW3-2 GBW3-3 GBW3-4 GBW3-5 GB6 GB10-1 GB10-2 Mean
Depth (m) 132.05 133.10 136.50 147.00 154.05 103.00 104.00 129.50 144.55 151.54 177.00 181.10 184.50 160.00 171.30 171.35
Drill hole DD042 DD042 DD042 DD042 DD042 DD043 DD043 DD036 DD042 DD042 DD042 DD042 DD042 DD042 DD043 DD043

Li 1.7 1.6 4.1 2.0 1.4 2.7 4.1 0.2 2.23 8.6 9.0 1.5 1.0 8.6 6.9 8.4 7.8 6.48
Be 0.48 0.44 0.73 0.63 0.46 0.30 0.31 0.40 0.47 0.15 0.18 0.58 0.31 0.41 0.49 3.06 2.07 0.91
Sc 0.4 0.3 2.1 0.2 0.3 0.4 1.4 0.2 0.66 14.6 13.4 2 2.4 7.5 2.3 7.9 6.6 7.09
V 6 <5 15 <5 <5 7 18 2 6 80 59 13 11 33 23 39 41 37.38
Cr 20 20 30 20 10 10 23 17 18.75 110 100 40 40 90 40 26 21 58.38
Co 0.6 0.8 3.0 0.9 1.0 1.6 2.7 1.5 1.51 176 62.4 42.3 11.9 133 4.8 4.4 3.1 54.74
Ni 2.4 2.5 5.4 2.3 3.2 4.3 4.1 2.5 3.34 289 96.6 41.6 9.3 124.5 7.6 10.2 5.9 73.09
Cu 1.4 0.7 5.5 0.7 2.3 0.5 1.7 1.1 1.74 2460 390 319 30 1170 34.8 95.8 39.0 567.33
Zn 24 17 46 20 19 15 67 24 29.00 275 137 341 279 188 54 53 49 172.00
Ga 0.9 0.9 5.0 0.8 0.9 2.3 7.1 1.3 2.40 10.6 8.3 3.3 3.0 7.6 8.2 7.2 7.1 6.91
Rb 0.3 0.2 0.9 0.2 0.2 1.0 4.3 0.3 0.93 5.7 1.4 0.7 0.5 3.0 19.2 32.1 32.4 11.88
Sr 9.5 12.1 9.5 9.7 9.6 10.0 39.4 12.7 14.06 13.7 5 11.7 15.7 33.2 20.1 285 532 114.55
Zr 5 4 14 2 2 9 27 5 8.50 1210 231 17 34 79 32 219 215 254.63
Nb 0.4 0.3 1.5 0.2 0.3 0.6 1.2 0.3 0.60 6.1 4.4 2.0 0.8 1.4 2.9 5.1 4.7 3.43
Cs 0.03 0.02 0.04 0.04 0.02 0.05 0.21 0.01 0.05 0.89 0.21 0.05 0.09 0.31 1.12 0.74 0.76 0.52
Ba 2.5 2.6 7.3 1.0 0.8 7.9 24.8 1.8 6.09 163.5 40.3 3.7 6.2 22.1 101.5 459 453 156.16
La 2.1 1.9 11.0 1.8 1.8 4.7 5.4 2.9 3.95 164.0 10.7 5.9 4.4 9.4 9.0 58.7 51.2 39.16
Ce 4.6 3.7 20.6 3.9 4.0 9.1 10.9 5.3 7.76 322 22.0 14.9 9.9 17.4 18.9 122.5 103.5 78.89
Pr 0.56 0.44 2.30 0.46 0.51 1.02 1.22 0.60 0.89 34.2 2.58 2.06 1.28 1.87 2.25 14.15 11.70 8.76
Nd 2.4 1.9 8.7 2.1 2.2 3.8 5.2 2.3 3.58 120.0 10.7 9.2 5.4 6.5 8.9 53.4 44.3 32.30
Sm 0.52 0.44 1.69 0.49 0.54 0.72 1.03 0.46 0.74 17.50 3.42 2.25 1.34 1.28 1.75 9.26 7.64 5.56
Eu 0.25 0.20 0.59 0.26 0.28 0.36 0.33 0.18 0.31 1.88 0.82 0.39 0.26 0.38 0.58 2.43 2.82 1.20
558

Gd 0.66 0.60 1.73 0.64 0.68 0.80 0.96 0.62 0.84 12.90 5.69 2.24 1.54 1.40 1.95 6.65 5.57 4.74
Tb 0.10 0.10 0.26 0.10 0.12 0.14 0.15 0.11 0.14 1.76 1.00 0.39 0.25 0.26 0.30 0.81 0.65 0.68
Dy 0.69 0.64 1.71 0.74 0.84 0.89 0.93 0.71 0.89 9.84 6.13 2.41 1.64 1.94 1.83 4.05 3.18 3.88
Y 5.1 4.9 9.7 5.7 5.8 6.7 6.4 7.0 6.41 41.2 30.2 15.3 11.5 13.1 11.6 21.6 17.7 20.28
Ho 0.17 0.15 0.35 0.17 0.19 0.22 0.22 0.17 0.21 1.81 1.21 0.56 0.38 0.47 0.39 0.78 0.60 0.78
Er 0.49 0.48 1.00 0.52 0.54 0.68 0.62 0.52 0.61 4.71 3.22 1.62 1.21 1.56 1.07 1.88 1.54 2.10
Tm 0.08 0.07 0.14 0.08 0.08 0.10 0.09 0.08 0.09 0.66 0.46 0.25 0.19 0.28 0.16 0.25 0.22 0.31
Yb 0.52 0.55 0.93 0.54 0.55 0.66 0.66 0.53 0.62 4.41 3.02 1.52 1.27 2.00 1.05 1.59 1.38 2.03
Lu 0.08 0.08 0.13 0.09 0.09 0.11 0.10 0.09 0.10 0.66 0.45 0.24 0.19 0.32 0.17 0.26 0.23 0.32
Hf <0.2 <0.2 0.4 <0.2 <0.2 0.3 0.8 0.2 0.21 29.9 5.8 0.4 0.2 2.1 0.8 5.5 5.2 6.24
Ta <0.1 <0.1 0.1 <0.1 <0.1 0.08 0.09 <0.05 0.03 0.2 0.2 0.1 <0.1 0.1 0.1 0.35 0.28 0.17
Au 0.010 0.005 <0.005 <0.005 <0.005 <0.005 <0.005 0.006 0.01 0.006 <0.005 <0.005 <0.005 <0.005 <0.005 <0.005 0.007 0.00
Pb 0.6 0.7 0.9 0.5 2.1 <0.5 1.4 <0.5 0.78 3.2 1.4 0.9 <0.5 5.2 1.5 5.9 10.2 3.54
Th 0.28 0.28 1.70 0.19 0.26 0.79 0.47 0.34 0.54 31.9 1.52 0.76 0.29 2.50 0.43 16.60 12.80 8.35
U 0.07 0.08 0.24 0.10 0.12 0.20 0.21 0.18 0.15 0.94 0.25 0.27 0.10 0.28 0.07 1.13 1.10 0.52
Th/U 4.00 3.50 7.08 1.90 2.17 3.95 2.24 1.89 3.34 33.94 6.08 2.81 2.90 8.93 6.14 14.69 11.64 16.14
ΣREE-Y 18.3 16.2 60.8 17.6 18.2 30.0 34.2 21.6 27.11 737.5 101.6 59.2 40.8 58.2 59.9 298.3 252.2 200.96
Y/Ho 30.00 32.67 27.71 33.53 30.53 30.45 29.09 41.18 31.89 22.76 24.96 27.32 30.26 27.87 29.74 27.69 29.50 27.51

Geoscience Frontiers 12 (2021) 549–572


(La/Yb)PAAS 0.23 0.20 0.68 0.19 0.19 0.41 0.47 0.32 0.34 2.14 0.20 0.22 0.20 0.27 0.49 2.13 2.14 0.97
(Sm/Yb)PAAS 0.40 0.32 0.72 0.36 0.39 0.43 0.62 0.34 0.45 1.57 0.45 0.59 0.42 0.25 0.66 2.31 2.19 1.08
(Pr/Yb)PAAS 0.27 0.20 0.62 0.21 0.23 0.39 0.46 0.28 0.33 1.93 0.21 0.34 0.25 0.23 0.53 2.22 2.11 1.08
(Ce/Ce*)PAAS 0.98 0.93 0.94 0.99 0.96 0.96 0.98 0.93 0.96 0.99 0.97 0.97 0.96 0.95 0.97 0.98 0.98 0.98
(Eu/Eu*)PAAS 2.02 1.77 1.62 2.17 2.04 2.10 1.52 1.48 1.84 0.58 0.82 0.76 0.83 1.22 1.47 1.47 2.08 1.08
(La/La*)PAAS 1.13 1.33 1.07 1.45 1.08 1.01 1.31 1.11 1.19 0.95 1.14 0.98 0.99 0.98 0.98 0.93 0.98 0.96
(Pr/Pr*)PAAS 0.99 0.97 1.01 0.94 1.00 1.02 0.95 1.01 0.99 1.02 0.99 1.02 1.02 1.03 1.02 1.03 1.02 1.02
(Y/Y*)PAAS 0.12 0.10 0.16 0.19 0.18 0.18 0.32 0.25 0.19 0.11 0.07 0.20 0.15 0.05 0.16 0.26 0.31 0.16

Note: post-Archean Australian shale (PAAS, McLennan, 1989). The REE-Y ratios are calculated as follow: (Ce/Ce*)PAAS ¼ CePAAS/(2PrPAAS  NdPAAS) (Bolhar et al., 2004), (Eu/Eu*)PAAS ¼ EuPAAS/(0.67SmPAAS þ 0.33TbPAAS)
(Bau and Dulski, 1996), (La/La*)PAAS ¼ LaPAAS/(3PrPAAS  2NdPAAS), (Pr/Pr*)PAAS ¼ 2PrPAAS/(CePAAS þ NdPAAS), (Y/Y*)PAAS ¼ 2YPAAS/(DyPAAS þ HoPAAS).
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 7. Spider diagrams of the amphibole BIFs (straight line) and the amphibole–pyrite BIFs (dash line) from the Gouap deposit displaying (a) the trace element
composition normalized to the Upper Continental Crust (Rudnick and Gao, 2003) and (b) the REE–Y data normalized to the Post Archean Australian Shale (PAAS;
Taylor and McLennan, 1985).

samples have Au content less than 0.005 ppm excepting three samples enrichment of transition metals such as Cu (30–2460 ppm), Zn (49–341
GB4-1 (0.01 ppm), GB4-2 (0.005 ppm) and GB9 (0.006 ppm). ppm), Ni (5.9–289 ppm), Co (3.1–176.0 ppm), Cr (21–110 ppm), V
The rare earth elements (ΣREE–Y) abundances in the amphibole BIFs (11–80 ppm), and Sc (2–14.6 pppm) (Table 2, Fig. 7a). These BIFs also
is distinctly moderate ranging from 16.2 to 60.8 ppm (average: 27.11 contain high and variable LILE contents including Sr (5–532 ppm) and Ba
ppm). The Post Archean Australian Shale (PAAS)-normalized REE–Y (3.7–453 ppm), except Rb and Cs which show much lower contents (<35
patterns of these BIF samples exhibit relative enrichment of HREE (Pr/ ppm). Zr has the highest contents (17–1210 ppm) for the HFSE in the
YbPAAS ¼ 0.20–0.62) with respect to LREE (La/YbPAAS ¼ 0.19–0.68) and amphibole–pyrite BIFs, reflecting the presence of zircon as observed in
enrichment of MREE (Sm/YbPAAS ¼ 0.32–0.72) with respect to HREE thin section (Fig. 5c–f). The other HFSE contents range from very low to
(Fig. 7b). The REE–Y patterns of all amphibole BIFs are also characterized moderate abundances (0.07–29.9 ppm), whereas the remainder trace
by: (i) slight positive La anomalies (La/La*PAAS ¼ 1.01–1.45), (ii) slight elements (Li, Be and Ga) have insignificant contents in these BIFs
negative to no Ce (Ce/Ce*PAAS ¼ 0.93–0.99) and Y anomalies (Y/Y*PAAS (Table 2, Fig. 7a). Most of these BIFs displays very low Au contents
¼ 0.94–1.02) anomalies, and (iii) pronounced positive Eu anomalies (<0.005 ppm) except for two samples GBW3-3 and GB10-2 with 0.006
((Eu/Eu*)PAAS ¼ 1.48–2.17). The amphibole BIFs show super-chondritic ppm and 0.007 ppm, respectively.
Y/Ho ratios varying between 27.71 and 41.18 (average: 31.89). When normalized to the PAAS (Taylor and McLennan, 1985), the
The amphibole–pyrite BIFs typically have high and variable trace amphibole–pyrite BIFs are obviously enriched in REE–Y (Fig. 7b). The
elements contents compared to the amphibole BIFs with notable PAAS-normalized REE–Y patterns for these BIFs (except samples

559
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Table 3
Silicon and oxygen isotopic compositions of the Gouap BIFs.
Drillhole Depth (m) Lithology Sample Mineral δ30SiNBS28 (‰) δ18OV-SMOW (‰) Δ18Omt-qtz (‰)

DD036 129.50 Amphibole BIF GB9 quartz 0.3 9.5 13


magnetite 3.5
DD043 103.00 Amphibole BIF GB11-1 quartz 1.3 7.5 9.6
magnetite 2.1
DD043 104.00 Amphibole BIF GB11-2 quartz 1.3 6.8 8.6
magnetite 1.8
DD043 105.00 Amphibole BIF GB11-3 quartz 1.3 7 9.2
magnetite 2.2
DD043 106.00 Amphibole BIF GB11-4 quartz 1 7.8
DD043 107.00 Amphibole BIF GB11-5 quartz 0.5 8.2
DD043 171.30 Amphibole–pyrite BIF GB10-1 quartz 0.8 10.6 12.9
magnetite 2.3
DD043 171.35 Amphibole–pyrite BIF GB10-2 quartz 0.9 9.2 12.2
magnetite 3
DD043 174.50 Amphibole–pyrite BIF GB10-3 magnetite 1.7

GBW3-1, GB6, GB10–1 and GB10-2) show depletion of LREE and MREE display variable Th and U contents ranging from 33.4 to 433.4 ppm and
relative to HREE, with (La/Yb)PAAS of 0.20–2.14 and (Sm/Yb)PAAS of from 149.9 to 949.2 ppm, respectively. Their distinct oscillatory zoning
0.25–2.31 (Fig. 7b). Their REE–Y patterns also display slight positive La and relatively high Th/U (0.12–0.77, average: 0.43, Table 3) indicate a
anomalies and slight negative Ce anomalies with (La/La*)PAAS and magmatic origin (e.g., Koschek, 1993; Belousova et al., 2002; Hoskin and
(Ce/Ce*)PAAS ratios 0.93–1.14 and 0.95–0.99, respectively. Negative to Schaltegger, 2003; Corfu et al., 2003; Li et al., 2019, 2020). This is
positive Eu anomalies ((Eu/Eu*)PAAS ¼ 0.58–2.08 and pronounced confirmed on the BSE images by the absence of alteration haloes or zircon
negative Y anomalies (Y/Y*)PAAS ¼ 0.05–0.31) is notable for most sam- overgrowths due to the overprint of hydrothermal events (Fig. 8a). The
ples. The amphibole–pyrite BIFs exhibit chondritic Y/Ho ratios ranging obtained ages from the most twenty-eight concordant (>95%) grain
from 22.76 to 30.26 (av: 27.51). spots range from 2377  24 Ma to 2929  22 Ma (Appendix S2), sug-
gesting Mesoarchean to Paleoproterozoic source provenance. The age’s
4.3. Isotopes composition distribution shows evidence of two main zircon populations. The first
population is composed of older detrital zircon grains, whose obtained
The isotopes data of the Gouap BIFs is presented in Tables 3 and 4. ages are concordant and plot near or on the concordia line and yield a
The δ30SiNBS28 values of the quartz range from 1.5‰ to 0.3‰ weighted mean 207Pb/206Pb age of 2619  34 Ma (MSWD ¼ 6.3, n ¼ 11;
(average: 0.95‰) and from 0.8‰ to 0.9‰ (average: 0.85‰) for Fig. 8b and c). The second and younger detrital zircon population (four
the amphibole BIFs and amphibole–pyrite BIFs, respectively (Table 3). grain spots) yielded a weighted average 207Pb/206Pb age of 2422  50
The quartz is characterized by δ18OV-SMOW of 6.8‰–9.5‰ (mean: 7.8‰) Ma (MSWD ¼ 1.9, n ¼ 4; Fig. 8b), suggesting early Paleoproterozoic
for the amphibole BIFs and of 9.2‰–10.6‰ (mean: 9.9‰) for the maximum depositional age (MDA) for the Gouap BIFs.
amphibole–pyrite BIFs (Table 3). The magnetite δ18OV-SMOW varies from Based on the CL images, most zircon grains in sample GBW3-1 are
3.5‰ to 1.8‰ (average: 2.4‰) and from 3‰ to 1.7‰ (average: euhedral short angular (56–84 μm) to rounded, whereas a few are short
2.3‰) for the amphibole BIFs and amphibole–pyrite BIFs, respectively to long prismatic (up to 130 μm) (Fig. 9a). Most zircon grains largely
(Table 3). The Δ18Omt-qtz values of the Gouap BIFs range from 8.6‰ to preserve magmatic oscillatory zoning at the core, with U-rich hydro-
13‰ (Table 3). The δ34SV-CDT values of the pyrite from the amphib- thermal replacement rims. On the BSE images, all the zircon grains of this
ole–pyrite BIFs vary between 1.1‰ and 1.8‰ with an average value of sample do not show any particular internal structure and appear homo-
1.37‰ (Table 4). geneous (Fig. 9a). These zircon grains commonly show low Th and
moderate U contents varying between 35.2–64.1 ppm and 64.9–670.8
4.4. LA-ICP-MS zircon U–Pb dating ppm, respectively. Most zircon grains display low Th/U ratio ranging
between 0.06 and 0.77 (Appendix S2) as well as high Nb, Hf and Ti
Zircon U–Pb age data for the two Gouap BIF samples (GB10-1 and contents, low LREE, and large positive Ce anomalies (not shown). This
GBW3-1) is reported in Appendix S2 and shown in Figs. 8 and 9. suggests that these zircon grains are late-magmatic to hydrothermal
Detrital zircon grains recovered from the sample GB10-1 are gener- zircon (e.g., Pettke et al., 2005; Schaltegger, 2007; Ayers and Peters,
ally euhedral to subhedral with size of 75–200 μm. CL images reveal that 2018; Jiang et al., 2019). Thirty-four analyzed spots yield apparent
207
most zircon grains are short to long prismatic with low to high lumi- Pb/206Pb ages varying between 2031 and 2194 Ma, with a weighted
nescence and distinct oscillatory zoning (Fig. 8a). These zircon grains average age of 2089  8.3 Ma (MSWD ¼ 0.80; Fig. 9b and c).

5. Discussion
Table 4
The δ34SV-CDT values (‰) of pyrite from the Gouap amphibole–pyrite BIFs from
drillhole DD043. 5.1. Contamination in BIFs and influence on REE patterns
34
Depth (m) Sample Mineral δ SV-CDT (‰)
5.1.1. Detrital input
171.30 GB10-1 pyrite 1.1 In most cases, iron formations are detritus-free (e.g., Dymek and
171.35 GB10-2 pyrite 1.5
Klein, 1988; Beukes and Klein, 1990; Bau and Dulski, 1996; Mloszewska
174.50 GB10-3 pyrite 1.4
174.75 GB10-4 pyrite 1.5 et al., 2012; Liu et al., 2014; Peng et al., 2018), but terrigenous
175.15 GB10-5 pyrite 1.1 contamination have been reported worldwide (e.g., Pecoits et al., 2009;
175.45 GB10-6 pyrite 1.8 Mloszewska et al., 2012; Gourcerol et al., 2016; Ghosh and Baidya, 2017;
175.90 GB10-7 pyrite 1.3
Aoki et al., 2018; Ganno et al., 2017; Soh Tamehe et al., 2018; Sun et al.,
177.10 GB10-8 pyrite 1.4
178.00 GB10-9 pyrite 1.3 2018; Hou et al., 2019). Some authors also report the association of
178.70 GB10-10 pyrite 1.3 chemically pure and impure BIFs within the same depositional

560
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 8. (a) Representative CL and BSE images with corresponding apparent U–Pb ages (Ma) of zircon from the sample GB10-1 (spot number in red and U–Pb age in
yellow). (b) Concordia plot of zircon U–Pb ages from the sample GB10-1. (c) LA-ICP-MS 207Pb/206Pb weighted mean age of zircon analyses from the sample GB10-1.

environment (Haugaard et al., 2013; Viehmann et al., 2015; Wang et al., BIFs are generally ascribed to admixture of clastic and volcanic material
2015; Barrote et al., 2017; Ndime et al., 2019). Overall, oxides (Al2O3 during BIF deposition (e.g., Lan et al., 2019), albeit the interaction be-
and TiO2), HFSE, LILE and transition metals are excellent monitor to tween later magmatic fluids and the BIFs may probably caused such REE
identify terrigenous contamination (e.g., Manikyamba et al., 1993; Arora abundance (Barrote et al., 2017). The Gouap BIFs display variable high
et al., 1995; Rao and Naqvi, 1995; Haugaard et al., 2013; Viehmann et al., ΣREE contents of 11.25–51.13 ppm (mean: 20.70 ppm) for the amphi-
2015; Wang et al., 2015). Zr, Rb, Sr, U and Th abundances in BIFs are bole BIFs and 29.25–696.33 ppm (mean: 180.69 ppm) for the amphib-
indicative of contamination from felsic crustal rocks, whereas the higher ole–pyrite BIFs. Fig. 11e–g shows that La, Sm and Yb contents are
contents of Cr, Ni, Co, Sc and V reflect input from mafic rocks (e.g., positively correlated with Zr contents, suggesting that the REE contents
Manikyamba et al., 1993; Arora et al., 1995; Rao and Naqvi, 1995). of the Gouap BIFs were influenced by detrital material.
The content of Al2O3þTiO2 in the Gouap amphibole BIFs is low (<0.8 Terrestrial materials such as felsic and basaltic rocks have a constant
wt.%) except for samples GB4-3 and GB11-2 that show elevated con- chondritic Y/Ho ratio of ~27, whereas seawater and associated chemical
centration of these elements (Table 1). Most of the Gouap amphib- sediments have superchondritic Y/Ho values (e.g., Bau and Dulski, 1999;
ole–pyrite BIFs (except samples GBW3-3 and GBW3-4) display high Bolhar et al., 2005). Little admixtures of clastic and volcanic components
Al2O3þTiO2 content ranging from 2.25 to 7.14 wt.% (Table 1). Addi- would depress seawater-like Y/Ho ratios of >44 (Bau and Dulski, 1996;
tionally, the Gouap BIFs show weak to strong positive correlation be- Bolhar et al., 2004). Most of the Gouap BIFs have higher Y/Ho ratio
tween Al2O3 and TiO2 (Fig. 10a), which may imply incorporation of (>27) and display a positive correlation between Zr content and Y/Ho
terrigenous components (Barrote et al., 2017, and references therein). ratios (Fig. 11h). These features suggest the involvement of detrital
This is also corroborated by slight to strong positive correlations between materials during the deposition of the Gouap BIFs.
Al2O3þTiO2, Zr and other trace elements (Figs. 10b–h and 11a–d). The
higher contents of Zr, Sr, Th, Cr, Ni, Co and Cu in the amphibole-pyrite 5.1.2. Hydrothermal and metasomatic alteration
BIFs compared to those of the amphibole BIFs are consistent with pres- Post-depositional processes such as hydrothermal and metasomatic
ence of felsic and mafic clastic components during the deposition of the alteration generally affect the trace metals and REE patterns in BIFs
Gouap BIFs (e.g., Manikyamba et al., 1993; Arora et al., 1995, and ref- particularly under high fluid/rock ratios (e.g., Hensler et al., 2015; Sil-
erences therein). veira Braga et al., 2015; Peng et al., 2018). At the Gouap area, the
Since Zr contents are low (ca. 2.2–10.2 ppm) in BIFs, previous studies occurrence of minerals such as epidote, chlorite and specially dissemi-
considered that the influence of contribution of clastic and volcanic nated pyrite in the amphibole-pyrite iron formations suggests secondary
components on REE contents is trivial (e.g., Bolhar et al., 2004; Frei and hydrothermal alteration. The chondrite-like signature (24 < Y/Ho < 34)
Polat, 2007; Friend et al., 2008). However, elevated ΣREE contents in of this rock also indicate the resetting of HREE content through the

561
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 9. (a) Representative CL and BSE images with corresponding apparent U–Pb ages (Ma) of zircon from the sample GBW3-1 (spot number in red and U–Pb age in
yellow). (b) Concordia plot of zircon U–Pb ages from the sample GBW3-1. (c) U–Pb age histogram displaying the results of zircon analyses from the sample GBW3-1.

interaction with hydrothermal fluids as already discussed by Hensler La and Eu anomalies (Bolhar et al., 2004) as depicted in their PAAS
et al. (2015) and Sampaio et al. (2018). Further evidence of hydrother- normalized REE–Y patterns (Fig. 13). Similar features have been
mal alteration is provided through the positive co-variation of the observed in other least contaminated BIFs (Fig. 13; Wang et al., 2015;
redox-sensitive elements such as Ce and Eu with ΣREE (Fig. 12a and b). Brando Soares et al., 2017; Soh Tamehe et al., 2018). However, the
studied BIFs show slight negative Y anomaly, which may be attributed to
5.2. Source of the Gouap BIFs slow rates of precipitation (Bau and Dulski, 1996). In Fig. 13, it is note-
worthy that the Gouap BIFs plot near the average compositions
The REE–Y signature of iron formations have been used as a proxy to low-temperature (<250  C) hydrothermal fluids (amphibole–pyrite BIFs)
identify the source of silica and iron (e.g., Bau and Dulski, 1996; Bolhar or lie between the average composition of Low-T hydrothermal fluids and
et al., 2004; Alexander et al., 2008; Planavsky et al., 2010; Silveira Braga South Pacific seawater (amphibole BIFs). On the other hand, the pro-
et al., 2015; Ganno et al., 2017; Soh Tamehe et al., 2018; Ndime et al., portion of seawater versus hydrothermal fluids is further emphasized in
2019). The conspicuous positive Eu anomalies in Archean and Paleo- Y/Ho vs. Eu/Sm and Sm/Yb vs. Eu/Sm diagram of Alexander et al.
proterozoic BIFs indicates the involvement of hydrothermal fluids from (2008). Plotted on both diagrams (Fig. 14a and b), most of the Gouap
deep-sea spreading centers to sub-oceanic waters (e.g., Bau and Dulski, BIFs cluster near seawater and low-temperature hydrothermal fluids
1996; Bolhar et al., 2004; Kato et al., 2006; Planavsky et al., 2010, and (0.1%) fields, and far away from high-temperature hydrothermal fluids
references therein). Strong positive anomalies characterize hydrothermal (5%). Therefore, we suggest that the REE–Y signatures of the Gouap BIFs
fluids of high-temperature (>250  C), while weak to no Eu anomaly are were inherited from low-temperature (<200  C) hydrothermal fluids.
typical of low-temperature (<200  C) fluids (Basta et al., 2011). The Using a combination of the above geochemical signatures, it can be
(Eu/Eu*)PAAS values of the Gouap amphibole BIFs vary from 1.48 to 2.17 summarized that the Gouap BIFs were precipitated from a mixture of
and amphibole–pyrite BIFs from 0.58 to 2.08 that inversely correlates with low-temperature hydrothermal fluids and seawater with relatively high
ΣREE (Fig. 12c), indicating that a secondary enrichment of these elements detrital contamination.
was responsible for the decreased anomaly. The positive correlation of Ni
vs. Cr, Co vs. Cr and Ni vs. Co (Fig. 10e–g), suggests a primary hydrothermal 5.3. Depositional environment and ocean redox chemistry
source for these elements (Dymek and Klein, 1988), though the influence of
detrital contaminant as well as hydrothermal and metasomatic alteration The Gross (1980) scheme has been widely used to classify the BIFs
should not be neglected as discussed in the previous section. into Algoma-type and Lake Superior-type based on their depositional
The Gouap BIFs exhibit distinctive (but less pronounced) seawater environment. Algoma-type BIFs are often interlayered with or strati-
signatures such as LREE-depletion, HREE-enrichment, and slight positive graphically associated with volcanic rocks in greenstone belts, whereas

562
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 10. Binary diagrams of elemental contents for the Gouap BIFs: (a) TiO2 vs. Al2O3, (b) Zr vs. (Al2O3 þ TiO2), (c) Rb vs. Sr, (d) Th vs. U, (e) Ni vs. Cr, (f) Co vs. Cr,
(g) Ni vs. Co, (h) Sc vs. V.

563
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 11. Zr variation diagrams of (a) Co, (b) Cr, (c) Ni, (d) (Sc), (e) La, (f) Sm, (g) Yb and (h) Y/Ho ratio for the Gouap BIFs.

564
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

a back-arc basin or continental marginal ocean environment, similar to


Superior-type BIFs (Ganno et al., 2017; Soh Tamehe et al., 2018). All the
Gouap BIFs (except five samples, see Table 2) are largely characterized by
low (Eu/Eu*)PAAS ratio < 1.8, that are typical of Superior-type BIFs
(Huston and Logan, 2004). Hence, we infer that the Gouap BIFs were
deposited in a region where sediments extended from continental shelf to
deep-water settings along craton margins as proposed for the Krivoy Rog
BIFs (Viehmann et al., 2015).
Most of the Gouap BIFs lack negative Ce anomalies in the (Pr/
Pr*)PAAS vs. (Ce/Ce*)PAAS diagram (Fig. 15; Bau and Dulski, 1996). This
suggests that the Gouap BIFs were precipitated in suboxic to relatively
anoxic seawater, albeit LREE-depletion relative to HREE of these BIFs in
PAAS-normalized diagram (Fig. 13) indicates deposition under oxidized
shallow seawater. The above interpretations imply that the Gouap basin
may have varied redox conditions and a strong redoxcline separating an
upper oxic water column from deeper waters that were suboxic to anoxic
(Planavsky et al., 2010).

5.4. Isotopic constrains of the Gouap BIFs

Although contaminated BIFs are poor seawater archive, their isotopes


compositions are particularly useful as proxies of paleo-environmental
conditions from where the BIFs were precipitated (e.g., Becker and
Clayton, 1976; Perry, 1983; and references therein). Si isotope compo-
sitions of quartz in BIFs have been applied on discriminating weathering
and hydrothermal sources of silica (e.g., Steinhoefel et al., 2010; Heck
et al., 2011), whereas O isotope studies of BIFs have been focused to
constrain paleo-temperatures as well as fluid-mineral interactions and
temperatures of equilibration or isotopic exchange (e.g., Gregory and
Criss, 1986; Valley, 2001; Hyslop et al., 2008). On the other hand, S
isotopic signatures of BIFs commonly distinguish biogenic and volcanic
sources of pyrite (e.g., Cameron, 1983; Bowins and Crocket, 1994).
Although our BIF samples were collected from an iron mineralized zone,
the minerals from the iron ore were not analyzed in this study.
In general, quartz with negative δ30SiNBS28 (<0‰) is thought to have
been derived from hydrothermal fluids (e.g., Ding et al., 1996; Andre
et al., 2006; Li et al., 2014), whereas quartz with positive δ30SiNBS28
(>0‰) point out the influence of a continental source or precipitation
from isotopically heavy seawater (Robert and Chaussidon, 2006; Van den
Boorn et al., 2007). The quartz in the Gouap BIFs is highly depleted in
30
Si with δ30SiNBS28 values ranging from 1.5‰ to 0.3‰ (mean:
0.95‰) and from 0.8‰ to 0.9‰ (mean: 0.85‰) for the amphibole
BIFs and amphibole–pyrite BIFs, respectively (Table 3). This implies that
the quartz of the Gouap BIFs was probably formed by precipitation from
hydrothermal fluids. Similar δ30SiNBS28 values are documented for the
quartz of the Huoqiu BIFs (Hou et al., 2017, 2019) and the Yuanjiacun
BIFs (Hou et al., 2014; Men et al., 2019) in the North China Craton.
Nonetheless, a terrigenous input of silica cannot be ruled out as a source
of the depleted silicon isotopes (δ30Si < 0) of quartz within the Gouap
BIFs (Wang et al., 2016).
Fig. 12. ΣREE variation plots of (a) Ce, (b) Eu and (c) (Eu/Eu*)PAAS for the The quartz within the Gouap BIFs have δ18OV-SMOW values of 6.8‰–
Gouap BIFs. 9.5‰ (mean: 7.8‰) and 9.2‰–10.6‰ (mean: 9.9‰) for the amphibole
BIFs and the amphibole–pyrite BIFs, respectively (Table 3). These
δ18OV-SMOW values falls between those for quartz within igneous rocks
the Lake Superior-type BIFs are typically deposited on stable continental and values for marine siliceous rocks, similar to siliceous rocks that
shelves and commonly in no association with volcanic rocks (Bekker formed by high-temperature hydrothermal sedimentation (Li and Jiang,
et al., 2010). 1995; Hou et al., 2014, 2019; Li et al., 2014). Although the Gouap BIFs
The Gouap BIFs, for instance, are mainly interbedded with meta- have been metamorphosed up to amphibolite facies conditions (Soh
sedimentary rocks including garnet gneisses, quartzites and staurolite Tamehe et al., 2019), this metamorphism and hydrothermal fluid ex-
micaschists (Soh Tamehe et al., 2019). Although these BIFs display some change are likely to have decreased rather than increased quartz
geochemical similarity with volcanic rocks in terms of transition metals δ18OV-SMOW values (Knauth and Lowe, 2003; Knauth, 2005; Robert and
such as Cr, Co, Ni and Cu, no metavolcanics or volcaniclastic sediments Chaussidon, 2006). This suggests that the real δ18OV-SMOW values of
have been reported in the Gouap deposit. This suggests that the deposi- quartz within the Gouap BIFs are expected to have been higher than the
tion of the Gouap BIFs is not linked to hydrothermal volcanic activity in obtained values of this study as proposed for the quartz from the Huoqiu
an island arc environment (e.g., Liu et al., 2014). Previous studies on BIFs BIFs (Hou et al., 2019) and the Yuanjiacun BIFs (Hou et al., 2014) in the
from the Nyong Group suggest that these BIFs were probably deposited in North China Craton.

565
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 13. Average PAAS-normalized REE–Y patterns of the Gouap BIFs compared with the average of seawater and hydrothermal fluids are from Bolhar et al. (2004)
and Bau and Dulski (1999), respectively. Data of Yuanjiacun BIFs (Wang et al., 2015), Biquinho BIFs (Brando Soares et al., 2017) and Kpwa-Atog Boga BIFs (Soh
Tamehe et al., 2018) are also shown for comparison.

Magnetite with δ18O values <þ0.9‰ indicates a precipitation from values of the Gouap BIFs are ranging from 8.6‰ to 13‰ (see Table 3).
low-temperature (400  C) hydrothermal processes, whereas magnetite Because the Gouap BIFs have been metamorphosed, it is useful to plot
with δ18O values >þ0.9‰ is interpreted to have originated from a δ18O values of quartz and magnetite, which often record discordant
magma or high-temperature magmatic fluids (~800–1000  C) (Jonsson temperatures in slowly cooled metamorphic rocks since these minerals
et al., 2013). The δ18O values of the magnetite from the Gouap BIFs are resistant to post-metamorphic exchange (Perry, 1983; Hoefs, 2009).
(3.5‰ to 1.8‰ for the amphibole BIFs and 3‰ to 1.7‰ for the In this plot which gives calculated temperature ranges for pair minerals,
amphibole–pyrite BIFs) fall within the range (4‰ to þ5‰) of weakly the quartz–magnetite of amphibole BIFs fall between 450 and 600  C;
metamorphosed Archean to Paleoproterozoic BIFs (Gutzmer et al., while the temperature indicated for this pair mineral of amphib-
2006). However, our data also overlap the overall δ18O isotope values ole–pyrite BIFs is  450  C (Fig. 16a). Conformable results were obtained
(9.5‰ to þ3.2‰) of some iron oxides formed during iron mineraliza- while using the plot of fractionation of Perry (1983) which also provides
tion in the Thabazimbi, Noamundi, Pic de Fon, Carajas, Hamersley, and temperature of metamorphism for pair minerals (Fig. 16b). Since the
Gorumahishani-Badampahar high-grade iron ore deposits (Gutzmer quartz and magnetite are the major components of the BIFs, we can es-
et al., 2006; Cope et al., 2008; E Silva et al., 2008, E Silva et al., 2013; timate that the Gouap BIFs were metamorphosed at 350–600  C, which
Thorne et al., 2009; Ghosh and Baidya, 2017). Because the Gouap BIFs corresponds well with their low- to medium-grade metamorphic
have undergone secondary alteration, we suggest that the observed δ18O conditions.
isotope signature of the magnetite from the Gouap BIFs might be Given that the presence of pyrite in the Gouap BIFs rather indicates
inherited from secondary hydrothermal alteration processes but without secondary hydrothermal alteration, their δ34S isotope signature may
significant iron enrichment. reveal an influence from hydrothermal fluids (E Silva et al., 2008, E Silva
Additionally, previous O isotope studies of BIFs have demonstrated et al., 2013). The δ34S values of the amphibole–pyrite BIFs ranging from
that quartz-magnetite oxygen isotope can be a useful geothermometer to 1.1‰ to 1.8‰ seem compatible with a chemical precipitation by hy-
constrain the thermal gradients from which the BIFs have been meta- drothermal fluids related to magmatic processes (Shanks, 2001). How-
morphosed (e.g., Becker and Clayton, 1976; Perry, 1983; Hyslop et al., ever, the range of δ34S data for the Gouap BIFs is closed to the δ34S values
2008; Hoefs, 2009; Li et al., 2013). It is noticeable that the Δ18Omt-qtz (between 2.84‰ and þ4.38‰, Table 4) obtained for twelve pyrite

566
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

providing a potentially precise depositional age for these rocks. We


conducted in-situ zircon dating of two amphibole–pyrite BIFs samples
(GB10-1 and GBW3-1) due to the petrographical evidence of this mineral
in thin section (see Fig. 6c, e) as well as to the detrital contamination as
interpreted from geochemical composition.
Our studies show that detrital zircon from sample GB10-1 have
oscillatory zoning (Fig. 8a) and high Th/U ratios (mean: 0.43, Appendix
S2), which are typical features of a magmatic origin. Thus, the
207
Pb/206Pb age of these zircon grains range from 2929 to 2378 Ma
indicating a Mesoarchean to Paleoproterozoic source. The youngest age
of detrital zircon defines the maximum depositional age (MDA) of the
Gouap BIFs at ca. 2422 Ma, thus during the early Paleoproterozoic time.
This age is similar to the previous SHRIMP U–Pb age of ~2423 Ma ob-
tained by Lerouge et al. (2006) on detrital zircon from metasediments in
the Nyong Group. The early Paleoproterozoic MDA is also consistent with
some geochemical characteristics of many Archean–Paleoproterozoic
transition BIFs (e.g., general trend of LREE depletion relative to HREE, a
weak positive Eu anomaly and no Ce anomalies) observed in the Gouap
BIFs (Planavsky et al., 2010; Fig. 15).
It is noteworthy that detrital zircons from sample GBW3-1 have no
oscillatory zoning and display multifaceted shapes with compositionally
homogenous texture on BSE images (Fig. 9a). Corfu et al. (2003) have
suggested that the development of homogenous texture in magmatic
zircon could be attributed to the modifications during late and
post-magmatic cooling. Additionally, the studied zircon grains display
magmatic cores surrounded by high-U light domains, probably formed by
metasomatism or recrystallization. Sample GBW3-1 is strongly hydro-
thermally altered as shown in the previous sections. This suggests that
the light regions surrounding the dark core may have formed by meta-
somatism. The analyzed zircon grains yield a 207Pb/206Pb age varying
between 2032 and 2194 Ma, with a weighted average age of 2089  8.3
Ma (MSWD ¼ 0.80). More recently, Loose and Schenk (2018) have
performed U–Pb SHRIMP dating of eclogite samples from the Nyong
Group and constrained the high-pressure metamorphic event at 2093 
45 Ma. This age, which is quite similar to that obtained in the Gouap
amphibole–pyrite BIFs, implies that the Nyong Group suffered a regional
metamorphism accompanied with hydrothermal activity during the
Paleoproterozoic, precisely at ca. 2.09 Ga.
Based on the above data, we propose that the Gouap BIFs were
deposited at ca. 2422 Ma and metamorphosed some ca. 330 Ma later,
during the Eburnean–Transamazonian orogeny.

5.6. Implication for the age of the Ntem complex BIFs and correlation with
the Brazilian IFs
Fig. 14. Composition of the Gouap BIFs plotted in diagram of (a) Y/Ho vs. Eu/
Sm and (b) Sm/Yb vs. Eu/Sm (after Alexander et al., 2008) displaying the In Cameroon, the BIF depositional age in the Ntem Group is relatively
deposition of the studied BIFs in seawater with contribution of <0.1% hydro- constrained through the pioneering studies made by Chombong and Suh
thermal fluids. (2013) and Ndime et al. (2019). Chombong and Suh (2013) have per-
formed SHRIMP U–Pb analyses on zircon from metavolcanics inter-
grains in the ~2.5 Ga Huoqiu BIFs from the North China Craton (Hou bedded with the Mbalam BIFs within the Ntem Group and reported an
et al., 2019). This suggests that the δ34S compositions of the Gouap BIFs age of 2883  20 Ma interpreted as the onset of BIF deposition. More
are akin to those of Superior-type BIF (Hou et al., 2019). recently, Ndime et al. (2019) proposed a Neoarchean age of 2679 Ma as
In summary, the different isotope data of the Gouap BIFs confirms the the minimum depositional age of the Nkout West BIFs based on the
influence of seawater as well as secondary hydrothermal alteration. whole-rock Pb–Pb applied on magnetite BIF. These results indicate a
regional period of BIF deposition from Mesoarchean to Neoarchean
within the Ntem Group.
5.5. Depositional age of the Gouap BIFs Although the ages of the BIF associated wallrocks in the Nyong Group
have not been precisely constrained yet, the detrital zircon ages of the
Wallrocks interlayered with the BIFs have been generally used to present study shed a new light on the geologic history of this region. The
constrain the depositional age (e.g., Zhang et al., 2011; Cabral et al., LA-ICP-MS zircon U–Pb dating results of sample GB10-1 range from 2929
2012; Chombong et al., 2013; Wang et al., 2015; Barrote et al., 2017; to 2378 Ma. The occurrence of the 2.9–2.3 Ga detrital zircon in the Gouap
Brando Soares et al., 2017; Lan et al., 2019; Tong et al., 2019; Silveira BIFs suggests the existence of Mesoarchean (~2.9 Ga) crust in the Nyong
Braga et al., 2019). However, some authors (e.g., Lai and Yang, 2012; Cen Group. This result is in good agreement with the ca. 2.92 Ga SIMS zircon
et al., 2012; Lan et al., 2014; Liu and Yang, 2015; Zhu et al., 2015; Kim U–Pb crystallization age of charnockites from the Ntem Complex (Li
et al., 2018; Liu et al., 2018; Nzepang Tankwa et al., 2020) have delivered et al., 2016), while Tchameni et al. (2000) have reported Pb–Pb zircon
direct dating of detrital zircon grains separated from the iron formations, evaporation age of 2666  2 Ma and 2687  3 Ma on high-K granitic

567
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Fig. 15. Composition of the Gouap BIFs plotted in a (Ce/Ce*)PAAS vs. (Pr/Pr*)PAAS diagram after Bau and Dulski (1996). Fields of late Paleoproterozoic BIFs (<2.0 Ga)
and Archean and early Paleoproterozoic BIFs (>2.4 Ga) are from Planavsky et al. (2010).

rocks from the same complex. The weighted mean 207Pb/206Pb detrital 6. Conclusion
age of 2619  34 Ma obtained in this study suggests an important Neo-
archean detrital component, probably of felsic composition (high-K The Gouap deposit is a recently discovered iron deposit located in the
granitoid), during the Gouap BIF deposition. Nyong Group of the Ntem Complex (southern Cameroon). This iron de-
Our direct LA-ICP-MS U–Pb zircon detrital ages has constrained the posit is characterized by amphibole and amphibole–pyrite BIFs inter-
MDA of the Gouap BIFs within the Nyong Group at ca. 2422 Ma, implying bedded with gneisses, micaschists and quartzites. Combining isotopic,
an early Paleoproterozoic age. This age is much younger than the Neo- geochemical and geochronological studies of the Gouap BIFs, we came to
archean age (2699  7 Ma; Chombong et al., 2017) reported in the the following conclusions:
magnetite gneisses from the Ngovayang iron ore deposit (see Fig. 2),
suggesting at least two different BIF units within the Nyong Group. The (1) The Gouap BIFs show variable high Fe2O3þSiO2 contents and
Gouap BIFs have also undergone regional hydrothermal activity and trace elements (e.g., Zr, Co, Cu, Ni and Sc), which suggests that
metamorphism during 2194–2032 Ma, which overlaps the age of the their chemical precipitation was influenced by variable terrige-
metamorphism (2100–2000 Ma) in the Nyong Group as reported by nous input and secondary hydrothermal alteration. Their REE–Y
previous authors (Toteu et al., 1994; Lerouge et al., 2006; Loose and patterns display LREE depletion and positive La anomalies, which
Schenk, 2018; Nzepang Tankwa et al., 2020). are inherited from seawater. Positive Eu anomalies further indi-
Comparable ages have been documented by previous geochronolog- cate the contribution of hydrothermal fluids.
ical studies for IF deposition in the Itabira Group (Cau^e and Gandarela (2) The isotopic data (δ30Si, δ18O and δ34S) suggest that the Gouap
Formations) from the Quadrilatero Ferrífero Mining District of the S~ao BIFs were formed by mixing of seawater and low-temperature
Francisco craton (Brazil). The age of the Cau^e IF is bracketed by the U–Pb hydrothermal fluids.
dates of ca. 2.6 Ga for the youngest detrital zircon from the underlying (3) The Gouap BIFs were probably deposited in an extensive basin
Moeda Formation (Machado et al., 1992, 1996; Hartmann et al., 2006; between an oceanic volcanic center and a continental margin
Martinez Dopico et al., 2017) and the Pb–Pb isochron date of 2419  19 setting, similar to paleogeographic settings proposed by Beukes
Ma for carbonates of the overlying Gandarela Formation (Babinski et al., and Gutzmer (2008) for a combined Hamersley-Transvaal Basin.
1995). These ages are comparable to the MDA (~2422 Ma) of the Gouap (4) Detrital zircon grains extracted from two BIF samples reveal that
BIFs. The ages of metamorphic monazite and titanite constrain the peak the Gouap BIFs were deposited at ca. 2422 Ma and experienced
of metamorphism at the late stage of the Transamazonian orogeny to ca. medium-grade metamorphism and metasomatism at ca. 2089 Ma
2.03–2.06 Ga (Machado et al., 1992; Noce et al., 1998; Aguilar et al., during the Eburnean–Transamazonian orogeny, like the Cau^e
2017). This latter age can be correlated with the metamorphic event of Formation at the SW-border of the S~ao Francisco protocraton.
the Gouap BIFs. Thus, the Nyong Group is part of the Palaeoproterozoic (5) The Cau^e and the Gouap IFs were probably deposited at the border
Transamazonian belt, which once extended from west-central Africa to of an adjoining Archean continental mass (protocratons) that were
northeastern Brazil, a consequence of the Eburnean–Transamazonian amalgamated in the Rhyacian period (Transamazonian Orogeny)
collision ca. 2100 Ma of the Congo and the S~ao Francisco cratons (Toteu and finally consolidated in the Brasiliano–Congo orogeny as the
et al., 1994; Feybesse et al., 1998; Penaye et al., 2004; Lerouge et al., S~ao Francisco-Congo craton.
2006; Nzepang Tankwa et al., 2020).

568
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Acknowledgements

This work forms part of the first author’s PhD dissertation at China
University of Mining and Technology (Xuzhou, China). The authors
thank the mineral exploration company G-Stones Resources Ltd., espe-
cially Mr. Bougne Dieudonne, for the permission to access unpublished
data and collect drillcore samples of the Gouap iron deposit. We also
thank all geologists and technicians of the mineral company for assis-
tance during fieldwork. Associate Editor Christopher Spencer and one
anonymous reviewer are gratefully acknowledged for their constructive
reviews and comments. This work was supported by the “Fundamental
Research Funds for the Chinese Central Universities” (Grant No.
2017CXNL03) and the Chinese Scholarship Council(Grant No.
2015120T19). Prof. Hou Kejun, Prof. Li Huan, Lu Yang, Quan Ruiping
and Jiang Weicheng are thanked for their help with LA-ICP-MS zircon
U–Pb analyses and suggestions for data processing. We are grateful to
L.C. Zhang, C.L. Wang, Z.D. Peng, X.X. Tong and A.P. Djoukouo Soh for
constructive discussion and helpful advice during an early presentation
of this study.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.


org/10.1016/j.gsf.2020.07.009.

References

Aguilar, C., Alkmim, F.F., Lana, C., Farina, F., 2017. Palaeoproterozoic assembly of the
S~ao Francisco craton, SE Brazil: new insights from U–Pb titanite and monazite dating.
Precambrian Res. 289, 95–115.
Alexander, B.W., Bau, M., Andersson, P., Dulski, P., 2008. Continentally-derived solutes in
shallow Archean seawater: rare earth element and Nd isotope evidence in iron
formation from the 2.9 Ga Pongola Supergroup, South Africa. Geochem. Cosmochim.
Acta 72, 378–394.
Anderson, T., 2002. Correction of common lead in U–Pb analyses that do not report 204Pb.
Chem. Geol. 192, 59–79.
Andre, L., Cardinal, D., Alleman, L.Y., Moorbath, S., 2006. Silicon isotopes in 3.8 Ga West
Greenland rocks as clues to the Eoarchaean supracrustal Si cycle. Earth Planet Sci.
Lett. 245, 162–173.
Aoki, S., Kabashima, C., Kato, Y., Hirata, T., Komiya, T., 2018. Influence of contamination
on banded iron formations in the Isua supracrustal belt, West Greenland: reevaluation
of the Eoarchean seawater compositions. Geosci. Front. 9, 1049–1072.
Arora, M., Govil, P.K., Charan, S.N., Uday Raj, B., Balaram, V., Manikyamba, C.,
Chatterjee, A.K., Naqvi, S.M., 1995. Geochemistry and origin of archean banded
iron–formation from the Bababudan schist belt, India. Econ. Geol. 90, 2040–2057.
Ayers, J.C., Peters, T.J., 2018. Zircon/fluid trace element partition coefficients measured
by recrystallization of Mud Tank zircon at 1.5 GPa and 800–1000  C. Geochem.
Cosmochim. Acta 223, 60–74.
Babinski, M., Chemale Jr., F., Van Schmus, W.R., 1995. The Pb/Pb age of the Minas
supergroup carbonate rocks, Quadrilatero Ferrífero, Brazil. Precambrian Res. 72,
235–245.
Barrote, V.R., Rosiere, C.A., Rolim, V.K., Santos, J.O.S., McNaughton, N.J., 2017. The
Proterozoic guanh~aes banded iron formations, southeastern border of the S~ao
Francisco Craton, Brazil: evidence of detrital contamination. Geol. Usp. Serie
Científica 17 (2), 303–324.
Basta, F.F., Maurice, A.E., Fontbote, L., Favarger, P., 2011. Petrology and geochemistry of
the banded iron formation (BIF) of Wadi Karim and Um Anab, Eastern desert, Egypt:
implications for the origin of Neoproterozoic BIF. Precambrian Res. 187, 277–292.
Bau, M., Dulski, P., 1996. Distribution of yttrium and rare-earth elements in the Penge
and Kuruman iron-formations, Transvaal Supergroup, South Africa. Precambrian Res.
Fig. 16. Diagrams depicting the metamorphic conditions for the Gouap BIFs: (a) 79, 37–55.
Bau, M., Dulski, P., 1999. Comparing yttrium and rare earths in hydrothermal fluids from
δ18O of quartz vs. δ18O of magnetite for metamorphic rocks (modified after the Mid-Atlantic Ridge: implications for Y and REE behaviour during near-vent
Kohn, 1999; Hoefs, 2009). (b) Plot of fractionation (Δ ¼ 1000 lnα) vs. Tem- mixing and for the Y/Ho ratio of Proterozoic seawater. Chem. Geol. 155, 77–90.
perature ( C) for different pair minerals including quartz-water (ΔQW), Becker, R.H., Clayton, R.N., 1976. Oxygen isotope study of Precambrian banded iron
siderite-water (ΔSW), calcite-water (ΔCW), magnetite-water (ΔMW), and formation, Hamersley Range, Western Australia. Geochem. Cosmochim. Acta 40,
quartz-magnetite (ΔQM) (modified after Perry, 1983). ΔQW, ΔMW, and ΔSW data 1153–1165.
Bekker, A., Slack, J.F., Planavsky, N., Krapez, B., Hofmann, A., Konhauser, K.O.,
are from Becker and Clayton (1976).
Rouxel, O.J., 2010. Iron formation: the sedimentary product of a complex interplay
among mantle, tectonic, oceanic and biospheric processes. Econ. Geol. 105, 467–508.
Belousova, E.A., Griffin, W.L., O’Reilly, S.Y., Fisher, N.I., 2002. Igneous zircon: trace
element composition as an indicator of source rock type. Contrib. Mineral. Petrol.
Declaration of competing interest
143, 602–622.
Beukes, N.J., Klein, C., 1990. Geochemistry and sedimentology of a facies transition from
The authors declare that they have no known competing financial microbanded to granular iron-formation in the early Proterozoic Transvaal
interests or personal relationships that could have appeared to influence Supergroup, South Africa. Precambrian Res. 47, 99–139.
Beukes, N.J., Gutzmer, J., 2008. Origin and paleoenvironmental significance of major
the work reported in this paper. iron formations at the Archean-Paleoproterozoic boundary. SEG Reviews 15, 5–47.

569
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Bolhar, R., Kamber, B.S., Moorbath, S., Fedo, C.M., Whitehouse, M.J., 2004. Ghosh, R., Baidya, T.K., 2017. Mesoarchean BIF and iron ores of the badampahar
Characterization of Early Archean chemical sediments by trace element signatures. greenstone belt, iron ore group, East Indian Shield. J. Asian Earth Sci. 150, 25–44.
Earth Planet Sci. Lett. 222, 43–60. Gourcerol, B., Thurston, P.C., Kontak, D.J., C^ ote-Mantha, O., Biczock, J., 2016.
Bolhar, R., Kamber, B., Moorbath, S., Collerson, K., 2005. Chemical characterization of Depositional setting of Algoma-type banded iron formation. Precambrian Res. 281,
Earth’s most ancient clastic metasediments from the Isua Greenstone Belt, Southern 47–79.
West Greenland. Geochem. Cosmochim. Acta 69, 1555–1573. Gregory, R.T., Criss, R.E., 1986. Isotopic exchange in open and closed systems. Rev.
Bouyo Houketchang, M., Penaye, J., Mouri, H., Toteu, S.F., 2019. Eclogite facies Mineral. Geochem. 16, 91–127.
metabasites from the Paleoproterozoic Nyong Group, SW Cameroon: mineralogical Gutzmer, J., Mukhopadhyay, J., Beukes, N.J., Pack, A., Hayashi, K., Sharp, Z.D., 2006.
evidence and implications for a high-pressure metamorphism related to a subduction Oxygen isotope composition of hematite and genesis of high-grade BIF-hosted iron
zone at the NW margin of the Archean Congo craton. J. Afr. Earth Sci. 149, 215–234. ores. Geol. Soc. Am. Mem. 198, 257–268.
Bowins, R.J., Crocket, J.H., 1994. Sulfur and carbon isotopes in Archean banded iron Hartmann, L.A., Endo, I., Suita, M.T.F., Santos, J.O.S., Frantz, J.C., Carneiro, M.A.,
formations: implications for sulfur sources. Chem. Geol. 111, 307–323. McNaughton, N.J., Barley, M.E., 2006. Provenance and age delimitation of
Brando Soares, M., Corr^ea Neto, A.V., Zeh, A., Cabral, A.R., Pereira, L.F., do Quadrilatero Ferrífero sandstones based on zircon U–Pb isotopes. J. S. Am. Earth Sci.
Prado, M.G.B., de Almeida, A.M., Manduca, L.G., da Silva, P.H.M., de Araújo 20, 273–285.
Mabub, R.O., Schlichta, T.M., 2017. Geology of the Pitangui greenstone belt, Minas Haugaard, R., Frei, R., Stendal, H., Konhauser, K., 2013. Petrology and geochemistry of
Gerais, Brazil: stratigraphy, geochronology and BIF geochemistry. Precambrian Res. the ~2.9 Ga Itilliarsuk banded iron formation and associated supracrustal rocks, West
291, 17–41. Greenland: source characteristics and depositional environment. Precambrian Res.
Cabral, A.R., Zeh, A., Koglin, N., Gomes Jr., A.A.S., Viana, D.J., Lehmann, B., 2012. Dating 229, 150–176.
the Itabira iron formation, Quadrilatero Ferrífero of Minas Gerais, Brazil, at 2.65 Ga: Heck, P.R., Huberty, J.M., Kita, N.T., Ushikubo, T., Kozdon, R., Valley, J.W., 2011. SIMS
depositional U–Pb age of zircon from a metavolcanic layer. Precambrian Res. 204– analyses of silicon and oxygen isotope ratios for quartz from Archean and
205, 40–45. Paleoproterozoic banded iron formations. Geochem. Cosmochim. Acta 75,
Cameron, E.M., 1983. Genesis of Proterozoic iron-formation: Sulphur isotope evidence. 5879–5891.
Geochem. Cosmochim. Acta 47, 1069–1074. Hensler, A.-S., Hagemann, S.G., Rosiere, C.A., Angerer, T., Gilbert, S., 2015.
Chombong, N.N., Suh, C.E., 2013. 2883 Ma commencement of BIF deposition at the Hydrothermal and metamorphic fluid-rock interaction associated with hypogene
northern edge of Congo craton, southern Cameroon: new zircon SHRIMP data “hard” iron ore mineralisation in the Quadrilatero Ferrífero, Brazil: implications from
constraint from metavolcanics. Episodes 36, 47–57. in-situ laser ablation ICP-MS iron oxide chemistry. Ore Geol. Rev. 69, 325–351.
Chombong, N.N., Suh, C.E., Ilouga, C.D.I., 2013. New detrital zircon U-Pb ages from BIF- Hoefs, J., 2009. Stable Isotope Geochemistry, sixth ed. Springer–Verlag, Berlin, p. 285.
related metasediments in the Ntem Complex (Congo craton) of southern Cameroon, Horstwood, M.S.A., Kosler, J., Gehrels, G.E., Jackson, S.E., McLean, N.M., Paton, C.,
West Africa. Nat. Sci. 5 (7), 835–847. Pearson, N.J., Sircombe, K., Sylvester, P.J., Vermeesch, P., Bowring, J.F.,
Chombong, N.N., Suh, C.E., Lehmann, B., Vishiti, A., Ilouga, D.C., Shemang, E.M., Condon, D.J., Schoene, B., 2016. Community-derived standards for LA-ICP-MS U-(Th-
Tantoh, B.S., Kedia, A.C., 2017. Host rock geochemistry, texture and chemical )Pb geochronology: uncertainty propagation, age interpretation and data reporting.
composition of magnetite in iron ore in the Neoarchaean Nyong unit in southern Geostand. Geoanal. Res. 40, 311–332.
Cameroon. B. Appl. Earth Sci. 126 (3), 129–145. Hoskin, P.W.O., Schaltegger, U., 2003. The composition of zircon and igneous and
Cen, Y., Peng, S.B., Kusky, T.M., Jiang, X.F., Wang, L., 2012. Granulite facies metamorphic petrogenesis. Rev. Mineral. Geochem. 53, 27–62.
metamorphic age and tectonic implications of BIFs from the Kongling group in the Hou, K.J., Li, Y.H., Gao, J.F., Liu, F., Qin, Y., 2014. Geochemistry and Si-O-Fe isotope
northern Huangling anticline. J. Earth Sci. 23 (5), 648–658. constraints on the origin of banded iron formations of the Yuanjiacun Formation,
Cope, I.L., Wilkinson, J.J., Boyce, A.J., Chapman, J.B., Herrington, R.J., Harris, C.J., 2008. Lvliang Group, Shanxi, China. Ore Geol. Rev. 57, 288–298.
Genesis of the Pic de Fon iron oxide deposit, Simandou Range, Republic of Guinea, Hou, K.J., Ma, X.M., Li, Y.H., Liu, F., Han, D., 2017. Chronology, geochemical, Si and Fe
West Africa. Rev. Econ. Geol. 15, 339–360. isotopic constraints on the origin of Huoqiu banded iron formation (BIF),
Corfu, F., Hanchar, J.M., Hoskin, P.W.O., Kinny, P., 2003. Atlas of zircon textures. Rev. southeastern margin of the North China Craton. Precambrian Res. 298, 351–364.
Mineral. Geochem. 53, 469–500. Hou, K.J., Ma, X.M., Li, Y.H., Liu, F., Han, D., 2019. Genesis of Huoqiu banded iron
Dickinson, W.R., Gehrels, G.E., 2009. Use of U–Pb ages of detrital zircons to infer formation (BIF), southeastern North China Craton, constraints from geochemical and
maximum depositional ages of strata: a test against a Colorado Plateau Mesozoic Hf-O-S isotopic characteristics. J. Geochem. Explor. 197, 60–69.
database. Earth Planet Sci. Lett. 228, 115–125. Huston, D.L., Logan, G.A., 2004. Barite, BIFs and bugs: evidence for the evolution of the
Ding, T., Jiang, S., Wan, D., Li, Y., Li, J., Liu, Z., 1996. Silicon Isotope Geochemistry. Earth’s early hydrosphere. Earth Planet Sci. Lett. 220, 41–55.
Geological Publishing House, Beijing (in Chinese). Hyslop, E., Valley, J., Johnson, C., Beard, B., 2008. The effects of metamorphism on O and
Dymek, R.F., Klein, C., 1988. Chemistry, petrology and origin by banded iron-formation Fe isotope compositions in the Biwabik Iron Formation northern Minnesota. Contrib.
lithologies from the 3800 Ma Isua supracrustal belt, West Greenland. Precambrian Mineral. Petrol. 155, 313–328.
Res. 39, 247–302. Ilouga, C.D.I., Suh, C.E., Ghogomu, R.T., 2013. Textures and rare earth elements
E Silva, R.C.F., Hagemann, S., Lobato, L.M., Rosiere, C.A., Banks, D.A., Davidson, G.J., composition of banded iron formations (BIF) at Njweng prospect, Mbalam iron ore
Vennemann, T., Hergt, J., 2013. Hydrothermal fluid processes and evolution of the district, southern Cameroon. Int. J. Geosci. 4, 146–165.
giant Serra Norte jaspilite-hosted iron ore deposits, Carajas mineral province, Brazil. Jackson, S.E., Pearson, N.J., Griffin, W.L., Belousova, E.A., 2004. The application of laser
Econ. Geol. 108, 739–779. ablation-inductively coupled plasma-mass spectrometry to in situ U–Pb zircon
E Silva, R.C.F., Lobato, L.M., Rosiere, C.A., Hagemann, S.G., Zuccheti, M., Baars, F.J., geochronology. Chem. Geol. 211, 47–69.
Morais, R., Andrade, I., 2008. A hydrothermal origin for the jaspilite-hosted giant Jiang, W.C., Li, H., Evans, N.J., Wu, J.H., 2019. Zircon records multiple magmatic-
Sierra Norte deposits in the Cajajas mineral province, Para state, Brazil. Rev. Econ. hydrothermal processes at the giant Shizhuyuan W–Sn–Mo–Bi polymetallic deposit,
Geol. 15, 255–290. South China. Ore Geol. Rev. 115, 103–160.
Feybesse, J.L., Johan, V., Triboulet, C., Guerrot, C., Mayaga-Mikolo, F., Bouchot, V., Eko Jonsson, E., Valentin, R.T., H€ ogdahl, K., Harri, C., Weis, F., Nilsson, K.P., Skelton, A.,
Ndong, J., 1998. The West Central African belt: a model of 2.5–2.0 Ga accretion and 2013. Magmatic origin of giant “Kiruna-type” apatite-ironoxide ores in central
two-phase orogenic evolution. Precambrian Res. 87, 161–216. Sweden. Sci. Rep. 3, 1644. https://doi.org/10.1038/srep01644.
Frei, R., Polat, A., 2007. Source heterogeneity for the major components of ~ 3.7 Ga Kato, Y., Yamaguchi, K.E., Ohmoto, H., 2006. Rare earth elements in Precambrian banded
banded iron formations (Isua Greenstone Belt, Western Greenland): tracing the iron formation: secular changes of Ce and Eu anomalies and evolution of atmosphere
nature of interacting water masses in BIF formation. Earth Planet Sci. Lett. 253, oxygen. Memoir of the Geological Society of America 198, 269–280.
66–281. Kim, C.S., Jang, Y.R., Samuel, V.O., Kwon, S.H., Park, J.-W., Yi, K.W., Choi, S.-G., 2018.
Friend, C.R.L., Nutman, A.P., Bennett, V.C., Norman, M.D., 2008. Seawater-like trace Petrogenesis, detrital zircon SHRIMP U-Pb geochronology, and tectonic implications
element signatures (REE þ Y) of Eoarchaean chemical sedimentary rocks from of the Upper Paleoproterozoic Seosan iron formation, western Gyeonggi Massif,
southern West Greenland, and their corruption during high-grade metamorphism. Korea. J. Asian Earth Sci. 157, 78–91.
Contrib. Mineral. Petrol. 155, 229–246. Knauth, L.P., 2005. Temperature and salinity history of the Precambrian ocean:
Ganno, S., Ngnotue, T., Kouankap Nono, G.D., Nzenti, J.P., Notsa, F.M., 2015. Petrology implications for the course of microbial evolution. Palaeogeogr. Palaeoclimatol.
and geochemistry of the banded iron-formations from Ntem complex greenstones Palaeoecol. 219, 53–69.
belt, Elom area, Southern Cameroon: implications for the origin and depositional Knauth, L.P., Lowe, D.R., 2003. High Archean climatic temperature inferred from oxygen
environment. Chem. Erde 75, 375–387. isotope geochemistry of cherts in the 3.5 Ga Swaziland Supergroup, South Africa.
Ganno, S., Moudioh, C., Nchare, N.A., Kouankap Nono, G.D., Nzenti, J.P., 2016. Geol. Soc. Am. Bull. 115, 566–580.
Geochemical fingerprint and iron ore potential of the siliceous itabirite from Kohn, M.J., 1999. Why most “dry” rocks should cool “wet”. Am. Mineral. 84, 570–580.
Palaeoproterozoic Nyong series, Zambi area, Southwestern Cameroon. Resour. Geol. Koschek, G., 1993. Origin and significance of the SEM cathodoluminescence from zircon.
66 (1), 71–80. J. Microsc. 171, 223–232.
Ganno, S., Njiosseu Tanko, E.L., Kouankap Nono, G.D., Djoukouo Soh, A., Moudioh, C., Kosler, J., Slama, J., Belousova, E., Corfu, F., Gehrels, G.E., Gerdes, A., Horstwood, M.S.A.,
Ngnotue, T., Nzenti, J.P., 2017. A mixed seawater and hydrothermal origin of Sircombe, K.N., Sylvester, P.J., Tiepolo, M., Whitehouse, M.J., Woodhead, J.D., 2013.
superior-type banded iron formation (BIF)-hosted Kouambo iron deposit, U-Pb detrital zircon analysis – results of an inter-laboratory comparison. Geostand.
Palaeoproterozoic Nyong series, Southwestern Cameroon: constraints from Geoanal. Res. 37, 243–259.
petrography and geochemistry. Ore Geol. Rev. 80, 860–875. Lai, X.D., Yang, X.Y., 2012. Characteristics of the banded iron formation (BIF) and its
Ganno, S., Tsozue, D., Kouankap, N.G.D., Tchouatcha, M.S., Ngnotue, T., Takam, G.R., zircon chronology in Yangzhuang, western Shandong Province. Acta Petrol. Sin. 28,
Nzenti, J.P., 2018. Geochemical constraints on the origin of banded iron formation- 3612–3622 (in Chinese with English abstract).
hosted iron ore from the Archaean Ntem Complex (Congo Craton) in the Meyomessi Lan, T.G., Fan, H.R., Santosh, M., Hu, F.F., Yang, K.F., Liu, Y.S., 2014. U–Pb zircon
area, Southern Cameroon. Resour. Geol. 68 (3), 287–302. chronology, geochemistry and isotopes of the Changyi banded iron formation in the
eastern Shandong Province: constraints on BIF genesis and implications for

570
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

Paleoproterozoic tectonic evolution of the North China Craton. Ore Geol. Rev. 56, formation-hosted Nkout iron ore deposit, north western Congo craton, Central West
472–486. Africa. J. Afr. Earth Sci. 148, 80–98.
Lan, C.Y., Yang, A.Y., Wang, C.L., Zhao, T.P., 2019. Geochemistry, U–Pb zircon Ndime, E.N., Ganno, S., Nzenti, J.P., 2019. Geochemistry and Pb–Pb geochronology of the
geochronology and Sm–Nd isotopes of the Xincai banded iron formation in the Neoarchean Nkout West metamorphosed banded iron formation, southern
southern margin of the North China Craton: implications on Neoarchean seawater Cameroon. Int. J. Earth Sci. 108 (5), 1551–1570.
compositions and solute sources. Precambrian Res. 326, 240–257. Nga Essomba, T.P., Ganno, S., Tanko Njiosseu, E.L., Ndema Mbongue, J.L., Kamguia
Lerouge, C., Cocherie, A., Toteu, S.F., Penaye, J., Milesi, J.P., Tchameni, R., Nsifa, N.E., Woguia, B., Soh Tamehe, L., Takodjou Wambo, J.D., Nzenti, J.P., 2020. Geochemical
Fanning, C.M., Deloule, E., 2006. Shrimp U/Pb zircon age evidence for constraints on the origin and tectonic setting of the serpentinized peridotites from the
paleoproterozoic sedimentation and 2.05Ga syntectonic plutonism in the Nyong Paleoproterozoic Nyong series, Eseka area, SW Cameroon. Acta Geochimica 39 (3),
Group, South-western Cameroon: consequences for the eburnean-transamazonian 404–422.
belt of NE Brasil and central Africa. J. Afr. Earth Sci. 44, 413–427. Noce, C.M., Machado, N., Teixeira, W., 1998. U–Pb geochronology of gneisses and
Li, H., Danisík, M., Zhou, Z.K., Jiang, W.C., Wu, J.H., 2020. Integrated U–Pb, Lu–Hf and granitoids in the Quadrilatero Ferrífero (southern S~ao Francisco craton): age
(U–Th)/He analysis of zircon from the Banxi Sb deposit and its implications for the constraints for Archean and Paleoproterozoic magmatism and metamorphism. Rev.
low-temperature mineralization in South China. Geoscience Frontiers 11(4), Bras. Geociencias 28, 95–102.
1323–1335. https://doi.org/10.1016/j.gsf.2020.01.004. Nzepang Tankwa, M., Ganno, S., Okunlola, A.O., Tanko, N.E.L., Soh Tamehe, L., Kamguia
Li, H., Zhou, Z.K., Evans, N.J., Kong, H., Wu, Q.H., Xi, X.S., 2019. Fluid–zircon interaction Woguia, B., Mbita, A.S.M., Nzenti, J.P., 2020. Petrogenesis and tectonic setting of the
during low-temperature hydrothermal processes: implications for the genesis of the Paleoproterozoic Kelle Bidjoka iron formations, Nyong group greenstone belts, south
Banxi antimony deposit, South China. Ore Geol. Rev. 114, 103137. western Cameroon: constraints from petrology, geochemistry and LA-ICP-MS zircon
Li, Y.H., Jiang, S., 1995. Silicon and oxygen isotope evidence for exhalative genesis of the U-Pb geochronology. Int. Geol. Rev. https://doi.org/10.1080/
Devonian Pb-Zn deposits in Feng-Tai region, Qinling Mountains, China. Chin. Sci. 00206814.2020.1793423.
Bull. 40, 87–88. Pecoits, E., Gingras, M.K., Barley, M.E., Kappler, A., Posth, N.R., Konhauser, K.O., 2009.
Li, W.Q., Huberty, J.M., Beard, B.L., Kita, N.T., Valley, J.W., Johnson, C.M., 2013a. Petrography and geochemistry of the Dales Gorges banded iron formations:
Contrasting behavior of oxygen and iron isotopes in banded iron formations revealed paragenetic sequence, source and implications for paleo-ocean chemistry.
by in situ isotopic analysis. Earth Planet Sci. Lett. 384, 132–143. Precambrian Res. 172, 163–187.
Li, Y.H., Hou, K.J., Wan, D.F., Zhang, Z.J., Yue, G.L., 2014. Precambrian banded iron Penaye, J., Toteu, S.F., Tchameni, R., Van Schmus, R.W., Tchakounte, J., Ganwa, A.A.,
formations in the North China Craton: silicon and oxygen isotopes and genetic Nsifa, E.N., Minyem, D., 2004. The 2.1 Ga west central African belt in Cameroon:
implications. Ore Geol. Rev. 57, 299–307. extension and evolution. J. Afr. Earth Sci. 39, 159–164.
Li, X.H., Chen, Y., Li, J., Yang, C., Ling, X.X., Tchouankoue, J.P., 2016. New isotopic Peng, Z.D., Wang, C.L., Tong, X.X., Zhang, L.C., Zhang, B.L., 2018. Element geochemistry
constraints on age and origin of Mesoarchean charnockite, trondhjemite and and neodymium isotope systematics of the Neoarchean banded iron formations in the
amphibolite in the Ntem Complex of NW Congo Craton, southern Cameroon. Qingyuan greenstone belt, North China Craton. Ore Geol. Rev. 102, 562–584.
Precambrian Res. 276, 14–23. Pettke, T., Audetat, A., Schaltegger, U., Heinrich, C.A., 2005. Magmatic- to hydrothermal
Li, X.H., Tang, G.Q., Gong, B., Yang, Y.H., Hou, K.J., Hu, Z.C., Li, Q.L., Li, W.X., 2013b. crystallization in the W–Sn mineralized Mole Granite (NSW, Australia): Part II:
Qinghu zircon: a working reference for microbeam analysis of U–Pb age and Hf and O Evolving zircon and thorite trace element chemistry. Chem. Geol. 220, 191–213.
isotopes. Chin. Sci. Bull. 58, 4647–4654. Perry Jr., E.C., 1983. Oxygen isotope geochemistry of iron formations. In: Trendall, A.F.,
Liu, Y.S., Gao, S., Hu, Z.C., Gao, C.G., Zong, K.Q., Wang, D.B., 2010. Continental and Morris, R.C. (Eds.), Iron Formations: Facts and Problems. Elsevier, Amsterdam,
oceanic crust recycling-induced melt-peridotite interactions in the Trans-North China pp. 359–371.
Orogen: U–Pb dating, Hf isotopes and trace elements in zircons from mantle Planavsky, N., Bekker, A., Rouxel, O.J., Kamber, B., Hofmann, A., Knudsen, A.,
xenoliths. J. Petrol. 51, 537–571. Lyons, T.W., 2010. Rare Earth Element and yttrium compositions of Archean and
Liu, L., Zhang, L.C., Dai, Y.P., 2014. Formation age and genesis of the banded iron Paleoproterozoic Fe formations revisited: new perspectives on the significance and
formations from the Guyang greenstone belt, western north China craton. Ore Geol. mechanisms of deposition. Geochem. Cosmochim. Acta 74, 6387–6405.
Rev. 63, 388–404. Quinn, M.Q., Hubbard, S.M., van Drecht, R., Guest, B., Matthews, W.A., Hadlari, T., 2016.
Liu, L., Yang, X.Y., 2015. Temporal, environmental and tectonic significance of the Record of orogenic cyclicity in the Alberta foreland basin, Canada Cordillera.
Huoqiu BIF, southeastern North China Craton: geochemical and geochronological Lithosphere 8, 317–332.
constraints. Precambrian Res. 261, 217–233. Rao, T.G., Naqvi, S.M., 1995. Geochemistry, depositional environment and tectonic
Liu, L., Zhang, H.S., Yang, X.Y., Li, Y.G., 2018. Age, origin and significance of the Wugang setting of the BIF of Late Archaean Chitradurga schist belt, India. Chem. Geol. 121,
BIF in the Taihua complex, southern North China Craton. Ore Geol. Rev. 95, 217–243.
880–898. Robert, F., Chaussidon, M., 2006. A palaeotemperature curve for the Precambrian oceans
Loose, D., Schenk, V., 2018. 2.09 Ga old eclogites in the Eburnian-Transamazonian based on silicon isotopes in cherts. Nature 443, 969–972.
orogen of southern Cameroon: significance for Palaeoproterozoic plate tectonics. Rudnick, R.L., Gao, S., 2003. The composition of the continental crust. In: Rudnick, R.L.
Precambrian Res. 304, 1–11. (Ed.), The Crust. Elsevier-Pergamon, Oxford, pp. 1–64.
Ludwig, K.R., 2003. User’s Manual for a Geochronological Toolkit for Microsoft Excel. Sampaio, G.M.S., Pufahl, P.K., Raye, U., Kyser, K.T., Abreu, A.T., Alkmim, A.R.,
Berkeley Geochronology Center Special Publication, pp. 1–75. Nalini Jr., H.A., 2018. Influence of weathering and hydrothermal alteration on the
Machado, N., Noce, C.M., Ladeira, E.A., de Oliveira, O.A.B., 1992. U-Pb geochronology of REE and δ56Fe composition of iron formation, Cau^e Formation, Iron Quadrangle,
the Archean magmatism and Proterozoic metamorphism in the Quadrilatero Brazil. Chem. Geol. 497, 27–40.
Ferrífero, southern S~ ao Francisco craton, Brazil. Geol. Soc. Am. Bull. 104, Schaltegger, U., 2007. Hydrothermal zircon. Elements 3, 51.
1221–1227. Silveira Braga, F.V., Rosiere, C.A., Queiroga, G.N., Rolim, V.K., Santos, J.O.S.,
Machado, N., Schrank, A., Noce, C.M., Gauthier, G., 1996. Ages of detrital zircon from McNaughton, N.J., 2015. The Statherian itabirite-bearing sequence from the Morro
Archean-Paleproterozoic sequences: implications for greenstone belt setting and Escuro ridge, Santa Maria de Itabira, Minas Gerais, Brazil. J. S. Am. Earth Sci. 58,
evolution of a Transamazonian foreland basin in Quadrilatero Ferrífero, southeast 33–53.
Brazil: evidence from zircon ages by laser ablation ICP-MS. Earth Planet Sci. Lett. Silveira Braga, F.V., Rosiere, C.A., Santos, J.O.S., Hagemann, S.G., McNaughton, N.J.,
141, 259–276. Salles, P.V., 2019. The Horto-Baratinha itabirite-hosted iron ore: a basal fragment of
Manikyamba, C., Balaram, V., Naqvi, S.M., 1993. Geochemical signatures of polygenetic the Espinhaço basin in the eastern S~ao Francisco Craton. J. S. Am. Earth Sci. 90,
origin of a banded-iron formation (BIF) of the Archean Sandur greenstone belt (schist 12–33.
belt) Karnataka nucleus, India. Precambrian Res. 61, 137–164. Shanks III, W.C., 2001. Stable isotopes in seafloor hydrothermal systems. Rev. Mineral.
Martin ez Dopico, M.I., Lana, C., Moreira, H.S., Cassino, L.F., Alkmim, F.F., 2017. U–Pb Geochem. 43, 469–525.
ages and Hf-isotope data of detrital zircons from the late Neoarchean- Slama, J., Kosler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M.,
Paleoproterozoic Minas basin, SE Brazil. Precambrian Res. 291, 143–161. Horstwood, M.S.A., Morris, G.A., Nasdala, L., Norberg, N., Schaltegger, U.,

Maurizot, P., Abessolo, A., Feybesse, J.L., Johan, L.P., 1986. Etude de prospection miniere Schoene, B., Tubrett, M.N., Whitehouse, M.J., 2008. Plesovice zircon — a new natural
du Sud-Ouest Cameroun. Synthese des travaux de 1978 a 1985. Rapport de BRGM 85, reference material for U-Pb and Hf isotopic microanalysis. Chem. Geol. 249 (1–2),
274 (in French). 1–35.
McLennan, S.M., 1989. Rare earth elements insedimentary rocks: influence of provenance Spencer, C.J., Kirkland, C.L., Taylor, R.J.M., 2016. Strategies towards statistically robust
and sedimentary processes. In: Lipin, B.R., McKay, G.A. (Eds.), Geochemistry and interpretations of in situ U–Pb zircon geochronology. Geoscience Front. 7, 581–589.
Mineralogy of Rare Earth Elements, vol. 21. Rev. Miner. Geochem., pp. 169–200 Soh Tamehe, L., 2020. Geology and Genesis of the Gouap Banded Iron Formation (BIF)-
Men, Y.K., Wang, E.D., Fu, J.F., Jia, S.S., You, X.W., He, Q.W., 2019. Geology and hosted Iron Deposit, South Cameroon. PhD thesis. China University of Mining and
geochemistry of the Yuanjiacun banded iron formation in Shanxi Province, China: Technology, p. 158 (in Chinese with English abstract).
constraints on the genesis. Geol. Mag. 156 (11), 1839–1862. Soh Tamehe, L., Nzepang, T.M., Wei, C.T., Ganno, S., Ngnotue, T., Kouankap, N.G.D.,
Mloszewska, A.M., Pecoits, E., Cates, N.L., Mojzsis, S.J., O’Neil, J., Robbins, L.J., Simon, S.J., Zhang, J.J., Nzenti, J.P., 2018. Geology and geochemical constrains on
Konhauser, K.O., 2012. The composition of Earth’s oldest iron formations: the the origin and depositional setting of Kpwa-Atog Boga banded iron formations (BIFs),
Nuvvuagittuq Supracrustal Belt (Quebec, Canada). Earth Planet Sci. Lett. 317–318, northwestern Congo craton, southern Cameroon. Ore Geol. Rev. 95, 620–638.
331–342. Soh Tamehe, L., Wei, C.T., Ganno, S., Simon, S.J., Kouankap, N.G.D., Nzenti, J.P.,
Ndema Mbongue, J.L., Ngnotue, T., Ngo Nlend, C.D., Nzenti, J.P., Suh, C.E., 2014. Origin Lemdjou, Y.B., Htun Lin, N., 2019. Geology of the Gouap iron deposit, Congo craton,
and evolution of the Formation of the Cameroon Nyong series in the western border southern Cameroon: implications for iron ore exploration. Ore Geol. Rev. 107,
of the Congo craton. Journal of Geosciences and Geomatics 2, 62–75. 1097–1128.
Ndime, E.N., Ganno, S., Soh Tamehe, L., Nzenti, J.P., 2018. Petrography, lithostratigraphy Steinhoefel, G., von Blanckenburg, F., Horn, I., Konhauser, K.O., Beukes, N.J.,
and major element geochemistry of Mesoarchean metamorphosed banded iron Gutzmer, J., 2010. Deciphering formation processes of banded iron formations from

571
L. Soh Tamehe et al. Geoscience Frontiers 12 (2021) 549–572

the Transvaal and the Hamersley successions by combined Si and Fe isotope analysis Toteu, S.M., Van Schmus, W.R., Penaye, J., Nyobe, J.B., 1994. U-Pb and Sm-Nd evidence
using UV femtosecond laser ablation. Geochem. Cosmochim. Acta 74, 2677–2696. of Eburnean and Pan African high grade metamorphism in Cratonic rock of southern
Suh, C.E., Cabral, A.R., Ndime, E., 2009. Geology and ore fabrics of the Nkout high-grade Cameroon. Precambrian Res. 67, 321–347.
haematite deposit, southern Cameroon. In: Angerer, T., Hagemann, S., Rosiere, C.A. Valley, J.W., 2001. Stable isotope thermometry at high temperatures. Rev. Mineral.
(Eds.), Smart Science for Exploration and Mining. Proceedings of the Tenth Biennial Geochem. 43, 365–413.
SGA Meeting. Society for Geology Applied to Mineral Deposit, Townsville, Australia, Van den Boorn, S., van Bergen, M., Nijman, W., Vroon, P., 2007. Dual role of seawater and
pp. 558–560. hydrothermal fluids in Early Archean chert formation: evidence from silicon isotopes.
Sun, J., Zhu, X.K., Li, Z.H., 2018. Confirmation and global significance of a large-scale Geology 35, 939–942.
early Neoproterozoic banded iron formation on Hainan Island, China. Precambrian Viehmann, S., Bau, M., Hoffmann, J.E., Münker, C., 2015. Geochemistry of the Krivoy Rog
Res. 307, 82–92. Banded Iron Formation, Ukraine, and the impact of peak episodes of increased global
Taylor, S.R., McLennan, S.M., 1985. The Continental Crust: its Composition and magmatic activity on the trace element composition of Precambrian seawater.
Evolution. Blackwell Publishing, Oxford, UK, p. 312. Precambrian Res. 270, 165–180.
Tchameni, R., Mezger, K., Nsifa, N.E., Pouclet, A., 2000. Neoarchaean evolution in the Wang, C.L., Konhauser, K.O., Zhang, L.C., 2015. Depositional environment of the
Congo craton: evidence from K rich granitoids of the Ntem Complex, southern Paleoproterozoic Yuanjiacun banded iron formation in Shanxi, China. Econ. Geol. 10,
Cameroon. J. Afr. Earth Sci. 30, 133–147. 1515–1539.
Teutsong, T., Bontognali, T.R.R., Ndjigui, P.-D., Vrijmoed, J.C., Teagle, D., Cooper, M., Wang, C.L., Konhauser, K.O., Zhang, L.C., Zhai, M.G., Li, W.J., 2016. Decoupled sources of
Vance, Derek, 2017. Petrography and geochemistry of the Mesoarchean Bikoula the 2.3–2.2 Ga Yuanjiacun banded iron formation: implications for the Nd cycle in
banded iron formation in the Ntem complex (Congo craton), Southern Cameroon: Earth’s early oceans. Precambrian Res. 280, 1–13.
implications for its origin. Ore Geol. Rev. 80, 267–288. Wiedenbeck, M., Alle, P., Corfu, F., Griffin, W., Meier, M., Oberli, F., von Quadt, A.,
Thorne, W., Hagemann, S., Vennemann, T., Oliver, N., 2009. Oxygen isotope Roddick, J.C., Spiegel, W., 1995. Three natural zircon standards for U-Th-Pb, Lu-Hf,
compositions of iron oxides from high-grade BIF-hosted iron ore deposits of the trace element and REE analyses. Geostand. Newsl. 19, 1–23.
Central Hamersley Province, Western Australia: constraints on the Evolution of Zhang, X.J., Zhang, L.C., Xiang, P., Wan, B., Pirajno, F., 2011. Zircon U–Pb age, Hf
hydrothermal fluids. Econ. Geol. 104 (7), 1019–1035. isotopes and geochemistry of Shuichang Algoma-type banded iron-formation, North
Trendall, 2002. The significance of iron-formation in the Precambrian stratigraphic China Craton: constraints on the ore-forming age and tectonic setting. Gondwana Res.
record. Spec. Publs int. Ass. Sediment. 33, 33–66. 20, 137–148.
Tong, X.X., Wang, C.L., Peng, Z.D., Huang, H., Zhang, L.C., Zhai, M.G., 2019. Zhu, M.T., Dai, Y.P., Zhang, L.C., Wang, C.L., Liu, L., 2015. Geochronology and
Geochemistry of meta-sedimentary rocks associated with the Neoarchean Dagushan geochemistry of the Nanfen iron deposit in the Anshan-Benxi area, North China
BIF in the Anshan-Benxi area, North China Craton: implications for their provenance Craton: implications for ~2.55 Ga crustal growth and the genesis of high-grade iron
and tectonic setting. Precambrian Res. 325, 172–191. ores. Precambrian Res. 260, 23–38.

572

You might also like