You are on page 1of 17

Journal of Geochemical Exploration 188 (2018) 240–256

Contents lists available at ScienceDirect

Journal of Geochemical Exploration


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

Garnierite mineralization from a serpentinite-derived lateritic regolith, T


Sulawesi Island, Indonesia: Mineralogy, geochemistry and link to hydrologic
flow regime

Wei Fua,b, , Yinmeng Zhanga, Chongjin Panga, Xiangwei Zenga, Xiaorong Huanga, Mengli Yanga,
Ya Shaoa, Henry Linb,c,d
a
Guangxi Key Laboratory of Hidden Metallic Ore Deposits Exploration, Guilin University of Technology, Guilin 541004, China
b
Department of Ecosystem Science and Management, The Pennsylvania State University, University Park, PA 16802, USA
c
State Key Laboratory of Loess and Quaternary Geology, Institute of Earth Environment, Chinese Academy of Sciences, Xi'an 710061, China
d
Shanxi Key Laboratory of Accelerator Mass Spectrometry Technology and Application, Xi'an AMS Center, Xi'an 710061, China

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

Keywords: Garnierite represents a significant nickel ore in many lateritic Ni deposits worldwide. To gain a better under-
Garnierite standing of its nature and origin, a well-developed garnierite-hosting transect from the Kolonodale area of East
Hydrologic flow regime Sulawesi, Indonesia, has been investigated using field geology, mineralogy and geochemical data. Garnierite
Preferential water flow occurs mainly in veins in the lower saprolite of a serpentinite-derived regolith. Mineralogically, it can be de-
Ni geochemistry
termined as an intimate mixture of Ni-rich serpentine-like (lizardite-Népouite) and talc-like (kerolite-pimelite)
Lateritic Ni deposit
phases. Results of EMP analyses indicate that Ni is preferentially enriched in the talc-like phases rather than the
serpentine-like phases. A sequential precipitation of mineral phases progressively enriched in Ni and Si to form
garnierite during weathering is suggested. The Ni-lizardite (2.63–8.49 wt% Ni) with elevated Fe (4.02–6.44 wt
%) may have been inherited from saprolite in a first instance and enriched in Ni by cation exchange processes.
Newly precipitated minerals are kerolite-pimelite (7.84–23.54 wt% Ni) and then followed by Ni-free quartz.
Minor amount of Népouite (23.47–28.51 wt% Ni) occur in laths along shrinkage cracks of previously formed
minerals, indicating a late stage paragenetic sequence. With emphasis on a hydrologic consideration, indicators
of a preferential flow regime are identified in the garnierite-hosting regolith, including: (i) non-uniform pattern
of the garnierite field occurrence, (ii) syn-weathering active nature of the garnierite-hosting structures, (iii) close
relationship between the garnierite occurrence and vertical FeeMn oxides pipes as well as FeeMn oxides pat-
ched areas, and (iv) specific physico-chemical property of the garnierite location with higher organic matter
concentrations but lower pH values compared to surroundings. It is proposed that the origin of garnierite is
closely linked to a preferential flow of oversaturated solutions through accessible conduits in the regolith.
Garnierite features as colloidal nature, high organic matter and low pH are key-parameters in metal transport
and deposition.

1. Introduction silicates (Brand et al., 1998; Butt and Cluzel, 2013). In terms of eco-
nomic significance, hydrous Mg-silicate ores represent the highest-
Lateritic Ni deposits are an important ore source, accounting for grade type (usually beyond 2 wt% and even > 5 wt% Ni occasionally)
over 60% of the world's nickel resource (Berger et al., 2011; Kuck, and, historically, accounts for about 32% of total lateritic Ni resources
2013). As a consequence of developments in lateritic ore processing with a mean grade of 1.44 wt%, whereas oxide ores represent the lar-
techniques (Mudd, 2010), the amount of Ni being extracted from la- gest reserve type, accounting for about 60% of total Ni with a mean
teritic ores has increased steadily in recent years, with lateritic Ni ore grade of about 1.0–1.6 wt% Ni (Butt and Cluzel, 2013).
now contributing over 40% of annual global nickel production (Kuck, Garnierite is a specific lateritic Ni ore of grass-green color, poor
2013). There are three general types of lateritic Ni ore, based on the crystallinity and fine-grained nature, and it is solely hosted in lateritic
dominant minerals hosting Ni: oxides, hydrous Mg silicates and clay Ni deposits of the hydrous Mg silicate type (Brand et al., 1998; Butt and


Corresponding author at: College of Earth Sciences, Guilin University of Technology, Jiangan Road No. 12, Guilin 541004, China.
E-mail address: fuwei@glut.edu.cn (W. Fu).

https://doi.org/10.1016/j.gexplo.2018.01.022
Received 3 November 2016; Received in revised form 22 December 2017; Accepted 27 January 2018
Available online 31 January 2018
0375-6742/ © 2018 Published by Elsevier B.V.
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Cluzel, 2013; Golightly, 2010). Due to containing a remarkably high Ni 2016; Fritsch et al., 2016; Wells et al., 2009), Central America and
concentration, garnierite represents a significant Ni ore type in many South America (e.g., Gleeson et al., 2004; Soler et al., 2008; Villanova-
lateritic Ni deposits, as reported from the Goro deposit (95.1 Mt, de-Benavent et al., 2014) as well as central Africa (e.g., Prendergast,
1.58 wt% Ni), New Caledonia (e.g., Wells et al., 2009), the Sorowako 2013), this case is marked by being associated with a tectonically active
deposit (153 Mt, 1.77 wt% Ni), Indonesia (e.g., Ilyasa et al., 2016; region at the circum-Pacific, a kind of serpentinite derived lateritic
Sufriadin et al., 2015), and the Falcondo deposit (46.5 Mt, 1.21 wt% regolith as well as a typical tropical rain forest climate. It could provide
Ni), Dominican Republic (e.g., Tauler et al., 2009; Villanova-de- a good example for understanding garnierite occurrences in Southeast
Benavent et al., 2014). It is called “green gold” by some mining geol- Asia. However, the nature and formation mechanism of the garnierite
ogists, and is considered a reliable guide to exploring for Ni ores in from the Kolonodale area have not been studied in detail.
laterites because of its striking color and distinctive field occurrence. The aim of this study is to: (i) reveal the property of garnierite in the
Moreover, garnierite is of primary importance for determining the study site from macro- to micro-scale by detailed field examination,
mining volumes that can be exploited due to its inhomogeneous oc- textural relationships, and new mineral and geochemical data; (ii)
currence (Cathelineau et al., 2016). However, as a poorly defined elucidate the difference between the garnierite and nearby weathered
weathering product, to gain a definitive identification of the nature of products by conducting a systematic examination along a garnierite-
garnierite remains a challenge. To address the nature of garnierite, hosting transect; and (iii) find evidence that could bridge the garnierite
extensive efforts have been directed at determining the classification occurrence and hydrologic flow regime in a profile scale, and propose a
and nomenclature of garnierite-forming minerals (e.g., Brindley and new model for garnierite formation by coupling Ni geochemical process
Hang, 1973; Brindley et al., 1977; Faust, 1966; Fritsch et al., 2016; with hydrologic process.
Manceau and Calas, 1985; Springer, 1974; Soler et al., 2008; Talovina
et al., 2008; Wells et al., 2009). The mineral species of garnierite has 2. Geological background and field occurrence
been identified as a series of Ni-bearing magnesium phyllosilicates,
including serpentine, talc, sepiolite, chlorite and smectite (Manceau The K-shaped Sulawesi Island lies at the convergence zone of three
and Calas, 1985). In most cases, almost all garnierite samples comprise tectonic plates: Eurasian, Pacific and Indian-Australian. Four lithotec-
a combination of two or more of the above mineral species. However, tonic belts are identified in this island (Fig. 1A): the West Sulawesi
there is still some garnierite that cannot be classified precisely due to its Volcano-Plutonic Arc Belt, the Central Sulawesi Metamorphic Belt, the
fine-grained nature, poor crystallinity and frequent occurrence as in- East Sulawesi Ophiolite Belt, and the Continental Fragments of Banggai-
timate mixtures of different mineral species (Villanova-de-Benavent Sula, Tukang Besi, and Buton (Mubroto et al., 1994). The study site, the
et al., 2014). Kolonodale area in Central Sulawesi Province, is located within the East
Extensive efforts were made to elucidate the origin of garnierite. Sulawesi Ophiolite Belt. This ophiolite belt represents a part of the
Most garnierite discovered in tropical belts is considered to be formed Circum Pacific Phanerozoic multiple ophiolite that were emplaced from
by intense weathering of ultramafic rocks, as in New Caledonia, the Cretaceous to the Miocene (Hall and Wilson, 2000). It extends
Indonesia, Dominican Republic and other tropical countries (e.g., 700 km from north to south and crops out over > 15,000 km2 on Su-
Cathelineau et al., 2016; Fritsch et al., 2016; Sufriadin et al., 2015; lawesi Island (Kadarusmana et al., 2004). Ultramafic bodies are widely
Tauler et al., 2009; Villanova-de-Benavent et al., 2014; Wells et al., exposed within this belt. They mainly consist of harzburgite and lher-
2009). Textural variations with poorly ordered materials suggest that zolite peridotite, and a large part of them has been serpentinized in
garnierites have been precipitated from colloidal suspensions at low- varying degrees. In the study area, ultramafic rocks have been re-
temperature conditions (Brindley et al., 1977). Gali et al. (2012) gave a gionally metamorphosed at moderate-high grade and, petrographically,
thermodynamic explanation for the temporal and spatial precipitation they can be determined as serpentinite (Fu et al., 2014) essentially
of different garnierite phases at standard temperatures and pressures. composed of serpentine and minor orthopyroxenes with trace clin-
Notably, garnierite is hosted in lateritic Ni deposits of the hydrous opyroxenes and spinels. There are two other lithological units in the
Mg silicate type (Brand et al., 1998; Butt and Cluzel, 2013), which in- surroundings of the ultramafic complex: alluvial and sedimentary la-
dicates that it is limited to those regions where ultramafic rock ex- custrine rocks of Quaternary and Cretaceous sedimentary rocks
posures, tropical climatic regime and active tectonic setting coexist (Kadarusmana et al., 2004). Faults, fractures, and joints, striking mainly
(Brand et al., 1998). Cluzel and Vigier (2008) and Villanova-de- in NNW direction, are largely observed in the serpentinite body due to
Benavent et al. (2014) suggest that a brittle syn-tectonic environment regional tectonic activities.
during ultramafic-rock weathering has played an important role in the Eocene-Miocene and Miocene-Recent plate movements have greatly
formation of garnierite. The origin of garnierite veins is also linked to influenced the geological features of Sulawesi Island. The inferred re-
the present-day topography and water table movement (Cathelineau gional uplift as a whole can be linked to underthrusting of oceanic
et al., 2016). However, Talovina et al. (2008) reported that some gar- crust, which resulted in the surface exposure of ultramafic complexes
nierites from the Urals differ significantly from their counterparts of with a large area from the central to north of the Island (Golightly,
weathering origin. There are two paragenetic mineral associations: 1979). It appears that these ultramafic rocks, including the serpentinite
hydrothermal association and exogenous association, and those hy- body in Kolondale area, have been exposed since the Miocene. The
drothermal associated garnierite-forming minerals (pecor- climate in the study area is typical of a humid tropical rainforest with
aite–chrysotile–quartz) were promoted by the contact-metasomatic high temperatures and abundant rainfall, which support lush vegetation
impact of local thermal field and fluids of diorite massif (Varlakov, over the entire region. Local geomorphology is characterized by a low-
1992), which suggest that the Uralian garnierites are of both hydro- moderate relief with gently sloping surfaces. Under these prolonged and
thermal and weathering origin (Talovina et al., 2008). Fritsch et al. pervasive weathering conditions the exposed serpentinite has been
(2016) proposed a new model for interpreting the formation of gar- subjected to strong chemical alteration, giving rise to a thick lateritic
nierite and accompanying minerals (deweylite), suggesting that some regolith. Such regolith hosts lateritic Ni deposit. Also, soils derived from
garnierite infillings in faulted peridotites of New Caledonia are prob- ultramafic bedrock have a number of extreme chemical properties,
ably of alteration closely linked to post-obduction tectonic activity. which influences the development of a specific ecosystem renowned for
Fu et al. (2014) first reported the discovery of garnierite in the high levels of endemism (e.g., plant species restricted to a limited
Kolonodale exploration area, Sulawesi Island, Indonesia, and its con- geographic area and the occurrence of nickel hyperaccumulators) (van
tribution as an important source of high-grade lateritic Ni ore to local der Ent et al., 2013).
mining. When compared to other typical garnierite occurrences In the study area, the thickness of serpentinite-derived lateritic re-
worldwide, such as those from New Caledonia (e.g., Cathelineau et al., golith is largely heterogeneous with the change of geomorphic site. It

241
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 1. (A) Simplified geological map of Sulawesi Island, Indonesia, modified after Kadarusmana et al. (2004), showing four lithotectonic belts identified in this island. The green colored
area indicates the distribution of the East Sulawesi Ophiolite complex, and the red dot shows the location of study site. (B) An ideal garnierite-hosting lateritic profile showing three
lithostratigraphic horizons defined at the Kolonodale mine area. (C) Sample locations along a garnierite-hosting transect. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

commonly ranges in 5–10 m at flat highland and gentle hillslopes, tens of centimeters (Fig. 2E). Also, garnierite-veins may have a variety
whereas at the base of hills the thickness may locally exceed 30 m. The of shapes at different locations, with a bifurcated, inflated, or com-
typical well-developed lateritic profile can be divided into three li- pound appearance. In some locations garnierite-veins with different
thostratigraphic horizons based on variations in color, texture and orientations intersect each other, showing a mesh-like structure. Be-
mineralogy (Fig. 1B). They are termed, from bottom to top, the sa- sides the vein-like garnierite, other occurrences include thin coatings on
prolite horizon, limonite horizon, and ferruginous cap (Fu et al., 2014). joints and fault planes or on the saprolite clast surfaces (Fig. 2F). Hand
These lithostratigraphic horizons may not be completely developed in specimen garnierites are heterogeneous and present a large number of
all regolith profiles, and locally part may be absent due to differential saprolite breccias with various sizes and shapes (Fig. 2D).
weathering or erosion processes.
Garnierite is present in joints and fractures at the lower part of the
3. Samples and methods
saprolite horizon (Fig. 2), where hard blocks of slightly weathered rock
occur below the soft saprolite and limonite. Locally, it may extend into
Garnierite and its nearby weathered products have been system-
the underlying harder bedrocks along fractured zone where high den-
atically sampled along a typical garnierite-hosting transect (Fig. 1C),
sity of joints or fractures occur. Most of garnierite occurs as fillings in
including four representative garnierite samples (G-1, G-2, G-3, and G-
fractures or joints with a jade-green color and vein-like structure, and
4) and four saprolite samples (S-1, S-2, S-3, and S-4) that occur adjacent
can be discriminated from the nearby earthy and grayish-green sapro-
to the garnierite vein at different distances. All samples were analyzed
lite. The width of vein-like garnierite is commonly on millimeter to
using multiple methods, including pH and organic matter measure-
centimeter scale and its length may extend from several centimeters to
ments, X-ray powder diffraction (XRD), optical and scanning electron

242
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 2. Macrophotographs of garnierite and related weathered materials in the Kolonodale lateritic profile. (A) Three lithostratigraphic horizons along the lateritic profile and some
specific ferruginous pipes; (B) A garnierite concentrated area in the lower part of the saprolite horizon; (C) Fe-oxides and Mn-oxides patches surrounding the garnierite veins; (D) Typical
hand specimen of garnierite with green color, massive or warty texture, and impurities from the saprolite. (E) A garnierite stringer-vein filling in the joint; (F) Garnierite occurring as thin
coatings on the surface of the saprolite clasts. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

microscopy (SEM-EDS), X-ray fluorescence (XRF), inductively coupled concentrations > 0.1 wt%. The detection limits for the major elements
plasma mass spectrometry (ICP-MS), and electron microprobe (EMP). are generally better than 30 mg/kg. For trace elements (Sc, Ti, V, Co,
The pH values were measured at a soil to water mass ratio of 1:2.5 Cu, Zn, Ge, Rb, Sr, Y, Zr, Nb, Cs, Ba, Hf, Ta, Pb, Th and U) analysis, the
using the Mettler Toledo acidity meter with a precision of ≤0.01. samples were first digested by an HNO3 + HF acid mixture in high-
Organic matter was determined by the dichromate oxidation method. pressure bombs, and then measured using a Perkin-Elmer Elan 6000
Briefly, 1 g of soil was oxidized at 150 °C for 1 h with 25 ml of a mix of ICP-MS with detection limits of about several ng/kg in solution, cor-
potassium dichromate and sulfuric acid. Then, a reverse titration of responding to a determination limit of about 10 μg/kg in solid samples
Cr2O72− ions was made using an acidified ferrous ammonium sulfate for trace elements. The uncertainties of the ICP-MS analyses are esti-
solution. mated to be better than 5% (relative) for most trace elements with
The XRD data were obtained using a Philips X' Pert MPD dif- concentrations > 10 mg/kg and ~10% for elements < 10 mg/kg.
fractometer, working on Cu target at 40 kV and 40 mA. The range of 2θ The EMP analysis was performed on polished sections to determine
scanning was from 5° to 80°. The scan step and step duration were re- the chemical composition of minerals using JXA-8230 electron mi-
spectively 0.05° and 3 s. Prior to the XRD analysis, the garnierite sam- croprobe with a wavelength dispersive system. The measurement con-
ples were carefully separated by handpicking in order to obtain the ditions were 20 kV accelerating voltage, 15 nA probe current, and 2 μm
most pure phase, and then ground in an agate mortar. beam diameter. Calibrations were performed using natural and syn-
Samples were cut into thin sections for petrographic examination thetic reference materials: forsterite (Mg, Fe, and Si), anorthite (Ca),
using a Leica DMEP optical microscope under the plane- and crossed- albite (Na, Al), chromite (Cr), phlogopite (K), rutile (Ti), NiO (Ni), CoO
polarized light. The natural break surface of samples were coated with (Co), and MnO (Mn). Back-scattered electron images and elemental X-
carbon and examined by electron microscopy. A Carl Zeiss ΣIGMA field ray maps of Fe, Mg, Ni were made at 15 keV and 10 nA.
emission scanning electron microscope (FSEM) coupled with a SDD The whole rock geochemical analyses were carried out in the Key
Inca X-MAX20 X-ray energy dispersive spectroscope (EDS) were used Laboratory of Geochronology and Geochemistry, Chinese Academy of
for the micro-morphology study on the individual mineral, working at Sciences, whereas the petrography, pH, organic matter, XRD, SEM-EDS,
15 keV and 10 nA. and EMP were all conducted in the Key Laboratory of the Guangxi
Prior to whole-rock geochemical analysis, the garnierite and sa- Hidden Metallic Ore Deposits Exploration at Guilin University of
prolite samples were crushed into small chips or fragments, and then Technology.
dried at 70 °C for 24 h. After drying, the samples were ground into
powder at 200 meshes using an agate mortar, and then the powder was
4. Results
heated to 700 °C to destroy organic material before analysis. Major
elements (Si, Al, Ca, Fe, K, Mg, Mn, Na, Ni and P) were measured using
4.1. Garnierite samples
a Rigaku X-ray fluorescence spectrometer (XRF). The accuracies of the
XRF analyses are estimated to be ~2% (relative) for elements oxides
4.1.1. pH, organic matter, and bulk geochemistry
with concentrations > 2 wt%, and ~5% for those present in
The values of pH, organic matter, and major and trace elements of

243
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Table 1
Whole rock analyses of the garnierite and saprolite samples in the Kolonodale lateritic profile, including organic matter (g/kg), major element (wt%) and trace element (mg/kg).

Type Garnierite Saprolite Bedrock

Sample K-1 K-2 K-3 K-4 S-1 S-2 S-3 S-4 B1-1

pH 65.4 69.8 66.7 63.2 68.0 72.9 71.9 70.8 –

Organic matter 26.7 30.1 32.3 22.3 19.6 20.1 17.5 20.7 –

Al 0.06 0.08 0.21 0.05 0.37 0.71 2.11 1.21 0.94


Ca 0.03 0.03 0.05 0.02 0.09 0.09 0.07 0.12 1.06
Fe 0.36 2.42 1.65 0.89 8.43 7.08 21.93 10.08 6.01
K 0.02 0.02 0.02 0.01 0.02 0.02 0.01 0.02 0.02
Mg 10.45 9.54 7.34 10.06 14.96 19.00 4.32 19.90 21.58
Mn < 0.01 0.06 0.02 0.06 0.11 0.14 0.40 0.45 0.11
Na < 0.07 0.30 < 0.07 0.36 0.27 0.18 0.07 0.25 0.14
P < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01
Si 25.50 28.16 30.14 27.30 19.78 17.65 18.50 16.80 19.64
Cr 0.03 0.16 0.09 0.16 0.54 0.36 1.71 0.49 0.29
Ni 15.58 9.07 10.78 10.80 4.85 2.58 1.39 1.79 0.20
L.O.I. 7.55 7.86 6.34 8.36 11.77 13.88 11.35 9.87 9.19
Sc 1.42 2.59 2.97 2.35 10.95 13.22 32.42 23.28 12.51
Ti 7.45 13.45 39.64 34.40 57.92 138.50 90.00 128.90 200.6
V 18.52 20.92 29.33 21.92 50.94 44.33 151.10 144.33 12.51
Co 24.75 75.85 29.89 55.81 200.10 144.80 517.50 104.10 200.60
Cu 3.77 4.42 4.00 7.15 19.72 18.25 48.12 28.27 55.56
Zn 179.90 248.00 27.73 184.00 418.00 58.95 165.30 58. 59 102.00
Ge 21.59 17.20 12.06 17.10 4.29 2.78 2.57 2.71 15.41
Rb 1.06 0.81 1.42 2.87 0.40 0.15 0.11 0.12 37.43
Sr 0.28 0.10 0.56 0.31 0.23 0.26 0.69 0.47 1.09
Y 1.64 1.18 0.31 1.17 0.79 0.94 2.75 0.39 1.25
Zr 0.26 0.04 0.56 0.14 0.18 0.10 0.33 0.20 0.47
Nb 0.06 0.14 0.09 0.51 0.05 0.03 0.07 0.01 1.33
Cs 0.28 0.29 0.21 0.92 0.08 0.03 0.04 0.03 0.18
Ba 1.10 1.82 2.57 2.83 8.03 0.70 17.39 2.66 0.16
Hf < 0.01 < 0.01 0.02 < 0.01 < 0.01 0.01 0.05 0.03 0.09
Ta < 0.01 0.02 < 0.01 0.02 < 0.01 < 0.01 0.15 < 0.01 4.22
Pb 0.49 2.52 1.17 2.55 0.18 0.33 0.21 0.35 0.02
Th 0.03 0.01 0.06 0.01 0.02 < 0.01 0.01 < 0.01 0.07
U 0.72 < 0.01 0.24 0.12 0.27 0.02 0.06 0.02 3.01

the garnierite samples are listed in Table 1. The overall pH values of the pimelite solid solution series defined by Brindley et al. (1979), re-
four garnierite samples fall in the range of 6.32 to 6.98, and the organic presenting a hydrated talc with a general formula of (Mg,
matter concentration varies from 22.3 to 32.3 g/kg. Compared to other Ni)3Si4O10(OH)2·nH2O. Similarly, the kerolite-pimelite series in gar-
regolith samples reported by previous work (Fu et al., 2014), the gar- nierite is assigned respectively at 0.913–1.062 nm in samples from
nierite samples contains a slightly higher organic matter concentration Soroako, Indonesia (Sufriadin et al., 2015) and at 0.94–1.07 nm in
than that of saprolite horizon samples (avg. 22.0 g/kg), but its pH va- samples from Goro, New Caledonia (Wells et al., 2009). According to
lues is relative lower than that of the saprolite horizon samples (6.98 to Springer (1974), there is a shift in index peak of talc-like phases from
8.03). 0.93 nm to 1.0 nm, which might be attributed to the effects of addi-
The four samples have similar geochemical compositions, namely, tional water molecules in its structure. Furthermore, the broad peak at
high Si (25.50–30.14 wt%), Mg (7.34–10.45 wt%), and Ni 1.0 nm indicates that the talc-like phase has a relatively low crystal-
(9.07–15.58 wt%), but low Al (0.05–0.21 wt%), Fe (0.36–2.42 wt%), Cr linity. The index peak of the serpentine-like phase appears at about
(0.03–0.16 wt%), Ca (0.02–0.05 wt%), and Mn (< 0.06 wt%). This 0.7 nm, showing at 0.734–0.738 nm in all samples. This peak corre-
implies that the garnierite is rich in Si, Mg, and Ni, but poor in Al, Fe, sponds to the serpentine-népouite solid solution series with a general
Cr, Ca, and Mn. Notably, the average Ni concentration is about 11.56 wt formula of (Mg, Ni)3Si2O5(OH)4 (Brindley and Hang, 1973). In the lit-
%, which is over 11 times higher than that of the cutoff grade of the erature, similar index peak of the serpentine-népouite series is reported
general lateritic Ni ore (1.0 wt% Ni), and 57 times higher than the Ni at 0.731–0.738 nm in samples from Soroako, Indonesia (Sufriadin et al.,
concentration of the bedrock (0.20 wt% Ni). For trace elements, most 2015) and at 0.73–0.743 nm in samples from Goro, New Caledonia
show a lower value in the garnierite than that in the bedrock, including (Wells et al., 2009). Additionally, the quartz is identified by its
Sc, Ti, V, Co, Cu, Ba, and Pb, except Zn and Ge. Chemical composition 0.336 nm strong reflection.
of the garnierite with low Co concentration (24.75–75.85 mg/kg) in- The talc-like and serpentine-like phases could not be separated by
dicates that such ore has low potential in Co utilization compared to hand picking although XRD analyses indicate that they occur together
oxide type lateritic Ni ore (e.g., Brand et al., 1998). in the same sample. Under the optical or electron microscope, some of
these two phases could be distinguished according to their mineral
morphologies and textures. The aggregates of the serpentine-like phases
4.1.2. Mineralogy and textures are grayish, pale-green, and yellowish-brown in color in crossed polars,
Results of XRD analyses (Fig. 3) indicate that the mineral compo- and show vein-like, mesh-like or booklets microtexture (Fig. 4A, B). The
sition of the garnierite samples is dominated by talc-like and serpen- SEM images show that the mineral crystals of the serpentine-like phase
tine-like phases, with minor amount of quartz. The index peak of the are mainly euhedral to sub-euhedral. They are sheet-like (Fig. 4C),
talc-like phase is detected at about 1 nm, presenting at 0.996–1.009 nm needle-like, and tubular-like in shape (Fig. 4D), and commonly
in four samples. Such an index peak matches well with the kerolite-

244
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 3. XRD patterns of representative garnierite (upper diagram) and saprolite (lower diagram) sample showing their mineral components. (Mineral abbreviations: Srp—serpentine,
Ker–kerolite, Qtz—quartz).

Fig. 4. Typical microtextures of the aggregate of the serpentine-like phases in the garnierite. (A) The booklets microtexture in optical microscope photographs; (B) The mesh-like
microtextures in optical microscope photographs; (C) The sheet-like shape in the SEM image; (D) The tubular-like shape in the SEM image; (E) The Ni-serpentines in the back-scattered
electron (BSE) image and their Ni concentrations (wt%); (F) The népouites in the BSE images occurring as laths along shrinkage cracks and their Ni concentrations. (Mineral ab-
breviations: Srp—serpentine, Nep—népouite, Qtz—quartz).

245
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 5. Typical microtextures of the aggregate of the talc-like phases in the garnierite. (A) Spherical and botryoidal microtextures in the optical microscope photographs; (B) The
brecciaed microtextures in the optical microscope photographs; (C-D) The morphology of the talc-like phases with irregular shape and nearly below 1–2 μm grain size for individual
particles in the scanning electron microscope (SEM) images; (E) The microtextures of the talc-like phases (Kerolite) in the back-scattered electron (BSE) images, showing oscillation
laminas within the aggregates, and their Ni (wt%) concentrations; (F) Ni concentrations in the oscillation laminas of the Kerolite. (Mineral abbreviations: Ker—kerolite, Qtz—quartz).

10–200 μm in size. In contrast, the aggregates of the talc-like phases are microtexture (Fig. 6A, B). Within these microtextures, serpentine-like
grayish to brown color in crossed polars, and show a banded, spherical phases always appear in the core of these microtextures whereas talc-
or botryoidal microtexture (Fig. 5A). They coexist with microcrystalline like phases occur outwards, indicating that these two garnierite-
quartz matrix, and the quartz is frequently as the inner infillings of the forming minerals may be formed at different stages.
aggregates of talc-like phases. Further toward the inner direction of the
quartz matrix, micro-pores are observed (Fig. 5A). This suggests that
quartz and talc-like phases have grown in voids. Moreover, some 4.1.3. Mineral chemistry
banded aggregates of the talc-like phases are fractured by quartz Microprobe chemical data of the serpentine-like and talc-like phases
veinlets and show brecciated clasts (Fig. 5B). According to the SEM in garnierite are summarized respectively in Table 2 and Table 3.
images, the individual talc-like phase is nearly below 1–2 μm in size and The serpentine-like phases mainly consist of Si, Mg, Ni, and Fe,
some particles maybe at the nano-scale. The crystals commonly show which account for > 98 wt% of total, along with minor Al and Ca
irregular sheet-like morphology and are closely aggregated (Fig. 5C, D). (> 0.1 wt%). Concentrations of other elements are commonly near or
Oscillation lamina is remarkable for the aggregate of the talc-like below detection limit (< 0.1 wt%). Nickel shows significant variation,
phases in the Back Scattered Electron (BSE) images (Fig. 5E, F), in- ranging from 2.63 wt% to 28.51 wt%. Notably, the Ni values are mainly
dicating that there were periodic changes in the chemistry of the talc- concentrated at two ranges at 2.63–8.49 wt% and 23.47–28.51 wt%.
like phases (Fig. 5F). Thus, the serpentine-like phases can be separated into two groups,
Notably, a large number of mm-μm scale clasts comprising serpen- namely the low-Ni type and the high-Ni type. The low-Ni type ser-
tine-like phases can be observed in the BSE images of the garnierite. pentine-like phases show a lower Ni/Fe ratio than that of the high-Ni
They are scattered randomly and coated by colloidal talc-like phases or type serpentine-like phases (avg. 0.96 vs avg. 10.1). Structural formulae
a matrix of microcrystalline quartz, showing a distinct enveloping of the serpentine-like phases are calculated on the basis of seven oxygen
atoms. Magnesium, Ni and Fe are allocated to the octahedral layer, and

246
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 6. Back-scattered electron images (A–B) de-


monstrate an enveloping microtexture within the
garnierite. The serpentine-like phases (Ni-serpen-
tine) present at the core of this microtexture have
low Ni concentrations (< 10 wt%), whereas the
talc-like phases (mainly kerolite) in the surround-
ings have higher Ni concentrations (> 10 wt%).
(Mineral abbreviations: Srp—serpentine,
Nep—népouite, Ker—kerolite).

Table 2
Representative EMP analysis (wt%) and structural formulae (in atoms per formula unit) of the serpentine-like phases in the garnierite samples.

Type Ni-lizardite Népouite

Sample k-1 k-2 k-2 k-2 k-2 k-2 k-2 k-2 k-3 k-3 k-4 k-1 k-1 k-1 k-2

Spot 1 2 3 4 5 6 7 8 9 10 11 1 2 3 4

Si 20.81 20.63 21.50 20.67 20.30 19.74 19.74 20.38 19.54 20.17 19.39 19.43 20.00 20.01 19.96
Ti 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.05 < 0.01 0.02 0.01 0.01 < 0.01 < 0.01 0.22 0.01
Al 0.45 0.52 0.28 0.15 0.21 0.02 0.01 0.03 0.01 0.18 0.03 0.01 0.04 0.02 0.02
Fe 6.44 5.61 5.24 5.57 5.78 4.85 4.46 5.16 4.07 6.35 4.02 2.68 2.03 4.13 2.25
Mg 14.36 13.90 17.59 17.40 16.67 16.55 17.21 14.23 18.39 17.02 18.58 5.75 5.80 6.48 7.06
Ni 4.26 7.69 5.85 2.63 2.78 4.40 4.57 8.49 4.63 3.17 4.98 26.42 28.51 23.47 24.36
Cr 0.08 0.24 0.08 0.04 0.04 0.02 0.02 0.05 < 0.01 0.02 0.02 < 0.01 0.08 0.07 < 0.01
Mn < 0.01 0.05 0.01 0.03 0.03 < 0.01 < 0.01 0.05 < 0.01 < 0.01 < 0.01 0.03 0.02 < 0.01 0.01
Co 0.02 < 0.01 < 0.01 0.02 0.04 < 0.01 < 0.01 < 0.01 < 0.01 0.03 0.03 0.01 < 0.01 < 0.01 < 0.01
Ca 0.17 0.23 0.14 0.13 0.16 0.12 0.08 0.14 0.06 0.17 0.17 0.11 0.11 0.16 0.19
K < 0.01 < 0.01 < 0.01 0.02 0.01 0.02 0.02 0.02 0.02 0.03 0.03 0.06 0.04 0.11 0.09
Na 0.03 0.02 0.01 0.04 0.03 0.01 0.06 0.01 0.04 0.07 0.07 < 0.01 < 0.01 0.02 0.04
Si 2.26 2.22 2.17 2.19 2.20 2.18 2.16 2.22 2.11 2.16 2.07 2.15 2.18 2.20 2.19
Ti < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01
Al 0.05 0.06 0.03 0.02 0.02 < 0.01 < 0.01 < 0.01 < 0.01 0.02 0.02 < 0.01 < 0.01 < 0.01 < 0.01
ΣTetr. 2.31 2.28 2.20 2.21 2.22 2.18 2.16 2.22 2.11 2.18 2.08 2.15 2.18 2.20 2.19
Fe 0.35 0.30 0.27 0.30 0.31 0.27 0.24 0.28 0.22 0.34 0.22 0.17 0.17 0.23 0.21
Mg 1.81 1.73 2.06 2.14 2.09 2.12 2.18 1.80 2.31 2.12 2.38 0.76 0.71 0.88 0.91
Ni 0.22 0.39 0.28 0.13 0.14 0.23 0.24 0.44 0.24 0.16 0.24 1.66 1.71 1.41 1.37
ΣOct. 2.38 2.45 2.61 2.57 2.56 2.62 2.67 2.53 2.76 2.62 2.83 2.59 2.60 2.52 2.49
Ca 0.01 0.02 0.01 0.01 0.01 0.01 0.01 0.01 < 0.01 0.01 0.01 0.01 0.01 0.01 0.01
K < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 0.01
Na < 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 < 0.01 0.01

Si with minor amount of Al occupies the tetrahedral sites. The calcu- for the talc-like phases. Considering the Al-free nature of the talc-like
lated structural formulae of the low-Ni and high-Ni type serpentine- phases, Si is the only cation in the tetrahedral site. Magnesium, Ni, and
like phases are respectively [(Mg(1.73–2.37)Ni(0.13–0.44)Fe(0.22–0.35)) Fe are dominant cations in the octahedral sites. The general structural
(Si(2.11–2.26)Al(0.00–0.06))O5(OH)4] and [Ni(1.41–1.71)(Mg(0.71–0.91) formulae is [(Mg(1.38–2.39)Ni(0.60–1.39)Fe(0.00–0.07))Si(3.94–4.31)O10(OH)2·
Fe(0.17–0.23))Si(2.15–2.20)O5(OH)4]. According to the 50% rule for solid (H2O)]. According to Nickel (1992), most of the talc-like phases can be
solution terminology (Nickel, 1992), the low-Ni type serpentine-like classified as kerolite (ideal formula is (Mg,Ni)3Si4O10(OH)2(H2O)) with
phases can be classified as Ni-lizardite (ideal formula is (Mg, Ni(apfu)/Mg(apfu) < 1. In addition, five grains analyzed yield a Ni
Ni)3Si2O5(OH)4) due to Ni(apfu)/Mg(apfu) < 1, whereas the high-Ni (apfu)/Mg(apfu) > 1, indicating that these can be grouped as pimelite
type can be grouped as népouite (ideal formula is (Ni, Mg)3Si2O5(OH)4) (ideal formula is (Ni, Mg)3Si4O10(OH)2(H2O)). Its calculated structural
because of Ni(apfu)/Mg(apfu) > 1 (Fig. 4E and F). Based on this formula is [(Mg(1.48–1.51)Ni(1.70–1.78)Fe(0.05–0.07))Si(3.76–3.85)O10(OH)2·
classification, Ni-lizardite is more abundant than népouite in the stu- (H2O)].
died garnierites. Notably, there are some deviations between the ideal structural
The talc-like phases also contain high concentrations of Si, Mg, and formula and our calculated cases, and the same issue is also reported by
Ni (Table 3). Both Fe and Al are detected as minor elements (< 1 wt%). several previous works (e.g., Tauler et al., 2009; Villanova-de-Benavent
Concentrations of Cr, Ti, Ca, Na, and K are near or even below the et al., 2014; Wells et al., 2009). In our case, the O/T ratios
detection line, indicating that these elements are negligible in the talc- ((Mg + Ni + Fe)/Si) of the Ni-lizardite and népouite are respectively of
like phase structure. Nickel concentration ranges from 7.84 wt% to 1.05–1.37 and 1.00–1.12, lower than the ideal value (O/T = 1.5) of the
23.54 wt%. There is also a large variation in the Ni concentration at the serpentine group minerals. Similarly, the O/T ratios of the kerolite and
microscopic scale. For example, the oscillation laminas of the talc-like pimelite are 0.53–0.85, deviating from the ideal value (O/T = 0.75) of
phases show fluctuations of up to 4 wt% Ni (Fig. 5F). the talc group minerals. Such deviations are also indicated by plotting
Structural formulae are calculated on the basis of 11 oxygen atoms in the Mg-(Ni + Fe)-Si ternary diagram (Fig. 7), showing that part of

247
W. Fu et al.

Table 3
Representative EMP analysis (wt%) and structural formulae (in atoms per formula unit) of the talc-like minerals in the garnierite samples.

Type Kerolite Pimelite

Sample K-1 K-1 K-1 K-1 K-1 K-1 K-1 K-1 K-2 K-2 K-2 K-2 K-3 K-3 K-4 K-4 K-1 K-1 K-2 K-3 K-3

Spot 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 1 2 3 4 5

Si 24.18 24.02 23.38 25.05 24.50 25.36 26.14 25.75 26.75 26.47 25.47 24.29 24.94 26.02 25.98 22.95 23.21 22.20 22.64 22.81 23.11
Ti < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.02 < 0.01 0.02 0.01 0.01 < 0.01 < 0.01 0.05 < 0.01 0.10 < 0.01 < 0.01 0.07 0.05 < 0.01 0.01
Al 0.01 0.01 0.02 0.01 0.03 < 0.01 0.02 0.03 < 0.01 0.01 0.01 0.02 0.01 < 0.01 0.01 0.07 0.05 0.03 0.05 0.03 0.08
Fe 0.65 0.51 0.43 0.55 0.51 0.04 0.02 0.02 < 0.01 0.65 0.02 0.63 0.47 0.56 0.46 0.78 0.80 0.90 0.79 0.65 0.88
Mg 12.59 10.84 10.35 10.51 12.42 7.75 9.94 11.23 10.82 11.36 10.76 11.27 9.73 10.34 11.05 11.89 7.43 8.17 7.57 6.81 7.78
Ni 7.84 11.92 12.18 15.84 8.13 11.66 9.53 7.89 15.64 13.63 15.43 11.87 14.14 15.40 11.73 11.42 21.62 23.54 21.50 22.46 21.44
Cr 0.01 0.03 0.01 0.05 0.01 0.03 0.07 0.02 0.01 0.03 < 0.01 < 0.01 0.02 0.05 < 0.01 0.01 0.02 < 0.01 0.02 0.03 0.03

248
Mn 0.02 < 0.01 < 0.01 0.05 < 0.01 < 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 0.01 0.02 < 0.01 < 0.01 < 0.01 < 0.01 0.02 < 0.01 < 0.01 < 0.01
Co < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.02 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01
Ca 0.01 0.03 0.01 0.01 0.03 0.05 0.04 0.01 0.12 0.17 0.11 0.14 0.13 0.09 0.06 0.23 0.21 0.11 0.24 0.17 0.25
K < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.02 0.01 0.02 0.05 < 0.01 < 0.01 0.03 0.03 0.02 0.01 0.13 0.02 0.17 0.06 0.02
Na < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.08 < 0.01 0.01 < 0.01 < 0.01 0.02 0.04 0.04 < 0.01 0.04 0.10 0.13 0.13 0.02
Si 3.97 3.97 3.97 3.92 3.99 4.27 4.20 4.14 4.00 3.98 3.95 3.94 4.01 3.99 4.04 3.84 3.85 3.76 3.81 3.82 3.84
Ti < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 0.01 0.01 < 0.01 < 0.01
Al < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 0.01 0.01 0.01 0.01 0.01
ΣTetr. 3.97 3.97 3.97 3.93 3.99 4.27 4.20 4.15 4.00 3.98 3.95 3.95 4.02 3.99 4.05 3.85 3.86 3.78 3.83 3.83 3.85
Fe 0.05 0.04 0.04 0.04 0.04 < 0.01 < 0.01 < 0.01 < 0.01 0.05 < 0.01 0.05 0.04 0.04 0.04 0.07 0.06 0.05 0.06 0.05 0.07
Mg 2.39 2.08 2.04 1.91 2.34 1.51 1.85 2.09 1.87 1.98 1.94 2.12 1.82 1.84 1.99 2.30 1.48 1.51 1.49 1.47 1.50
Ni 0.61 0.94 0.99 1.18 0.63 0.94 0.73 0.60 1.11 0.98 1.14 0.92 1.09 1.13 0.87 0.91 1.71 1.78 1.70 1.75 1.70
ΣOct. 3.06 3.06 3.06 3.15 3.02 2.45 2.59 2.70 2.99 3.01 3.08 3.09 2.94 3.01 2.90 3.28 3.25 3.34 3.25 3.27 3.27
Ca < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 0.01 0.02 0.01 0.02 0.01 0.01 0.01 0.03 0.02 0.03 0.03 0.02 0.03
K < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 0.01 0.02 0.01 < 0.01
Na < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 < 0.01 0.02 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 0.01 0.01 < 0.01 0.07 0.08 0.05 0.03 < 0.01
Journal of Geochemical Exploration 188 (2018) 240–256
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 7. Mg-(Ni + Fe)-Si ternary diagram of the Ni-bearing minerals in the garnierite samples, plotted as calculated in atoms per formula unit. (Mineral abbreviations: Sep = sepiolite,
Fal = falcondoite, Ker = kerolite, Pim = pimelite, Krp = karpinskite, Ni-Krp = Ni-karpinskite, Liz = lizardite, Nep = népouite).

the data plot outside the serpentine-series line and the talc-series line, spatial distribution Ni, Mg, and Fe within serpentine-like phases.
with a drifting toward more Si-rich phases. These results show that A selected enveloping microtexture, formed both by the serpentine-
excess Si is present in our calculated structural formula of the gar- like phases at the core and the surrounding talc-like phases, has been
nierite-forming minerals. A possible explanation may be a true excess of scanned for element mapping (Fig. 8D). Elements Ni, Mg, and Fe show a
Si in the octahedral site or some contaminations of microscopic silica discrete distribution feather on this map. Depleted Ni micro-areas occur
rods in intimate association with the talc-like and serpentine-like mi- predominantly in the position of the serpentine-like phases at the core,
nerals (Tauler et al., 2009; Wells et al., 2009). Also, the Si excess could whereas elevated Ni micro-areas occur exactly in the position of the
be explained by the existence of other phases with intermediate com- talc-like phases in the surroundings. However, the serpentine-like
positions between Ni-talc and serpentine, as karpinskite (T/O = 1) phases at the core host higher Fe concentration than the surrounding
(e.g., Villanova-de-Benavent et al., 2014). However, the existence of talc-like phases.
karpinskite as a mineral species is debatable and it is not accepted by
the International Mineralogical Association (IMA).
4.2. Saprolite samples

4.1.4. X-ray element mapping The values of pH, organic matter, and major and trace elements of
In order to examine the relationships between mineral composition, saprolite samples are listed in Table 1. The saprolite samples have
microtexture, and distribution of elements in our garnierite samples, higher pH value (6.80–7.29, with an average of 7.09) but lower organic
three garnierite-forming elements Mg, Ni, and Fe have been mapped in matter concentration (17.5–20.7 g/kg, with an average of 19.5 g/kg)
selected micro-areas. Element mapping for the aggregate of the talc-like compared with the above described garnierite samples. The con-
phases with oscillatory lamina and with botryoidal microtexture is re- centration of Ni is 2.66 wt% on average, which is about ¼ of the con-
spectively presented in Fig. 8A and B. Strong periodical changes of Ni centration recorded for the garnierites (avg. 11.56 wt%). Also, higher
and Mg concentrations are present across the oscillatory laminae within Mg (4.32–19.90 wt%), Fe (7.08–21.93 wt%), and Al (0.37–2.11 wt%)
the talc-like phases, showing a converse-related distribution patterns concentrations are found in the saprolite samples compared to those of
between Ni and Mg. These periodical changes in elemental distributions the garnierite. Additionally, there are considerable differences in che-
within the banded microtexture may be attributed to the variations in mical composition between the four saprolite samples, especially in Ni
the physico-chemical conditions under the aqueous environment during concentrations. For example, the S-1 sample contains higher Ni con-
the formation of the garnierite (Tauler et al., 2009; Villanova-de- centration (4.85 wt% Ni) than that of the S-2 (2.58 wt% Ni), S-3
Benavent et al., 2014). (1.39 wt% Ni), and S-4 sample (1.79 wt% Ni), indicating a roughly Ni
Element mapping for the aggregate of the serpentine-like phases is decreasing trend with the saprolite samples further away from the
shown in Fig. 8C. It is a mesh-like microtexture surrounded by a quartz garnierite location (Fig. 1C).
veinlet. There is remarkable heterogeneity of the element distribution The mineralogy of the saprolite samples mainly comprises serpen-
within this microtexture, especially for Mg and Fe, and a lower extent tine, along with a small amount of chromite (Figs. 3 and 9). The ser-
for Ni. This observation indicates a varying elemental composition pentine minerals are colorless to pale in color and tabular or vein-like in
among those differently-oriented serpentine veins. This result is con- shape, and commonly show mesh-like texture (Fig. 9A, B), which is
sistent with the EMP data. As demonstrated in Fig. 8C, many micro- similar to the characteristics of serpentine in the bedrock (Fu et al.,
areas with higher-Ni concentration commonly have lower Mg and Fe 2014). Chromite grains are present among the serpentine veins
concentrations, suggesting that there is an inverse correlation in the (Fig. 9A, C). Microprobe chemical data (Table 4) show that the

249
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 8. Mapping of the Fe, Mg, and Ni concentrations for selected areas within the garnierite. (A) Element distribution maps of the aggregate of the talc-like phases with a oscillatory
banded microtexture; (B) Element distribution maps of the aggregate of the talc-like phases with a botryoidal microtexture; (C) Elemental distribution maps for the aggregate of the
serpentine-like phases with a typical mesh-like microtextures; (D) Element distribution maps for a enveloping microtexture formed jointly by the serpentine-like phases at the core and the
talc-like phases in the surroundings. (Mineral abbreviations: Srp—serpentine, Nep—népouite, Ker—kerolite).

serpentines from the saprolite samples have a large variation in Si 5. Discussions


(18.04–20.43 wt%), Mg (13.82–20.96 wt%), Fe (3.98–7.96 wt%), and
Ni (1.18–4.00 wt%). Obviously, their Ni concentrations are much lower 5.1. Nature of the garnierite at Kolonodale Ni deposit
than those of the serpentine-like phases from the garnierite samples.
The structural formula of these serpentines have been calculated as: The mineralogical and geochemical data presented above indicate
[(Mg(1.81–2.55)Ni(0.05–0.22)Fe(0.21–0.45))(Si(2.01–2.23)Al(0.02–0.12))O5(OH)4], that the garnierite from Kolonodale Ni deposit is a binary mineral
which can be classified as Ni-lizardite. mixture composed of serpentine-like and talc-like phases. In a global
comparison, the data are analogous to those reported for garnierites
from the Goro deposit in New Caledonia (Wells et al., 2009) and the
Loma de Hierro deposit in Venezuela (Soler et al., 2008). Contrastingly,

Fig. 9. (A) Optical microscope photographs of the saprolite dominated by serpentine and a small amount of chromite; (B) Scanning electron microscope (SEM) images illustrating the
mess texture of the saprolite, inherited from the parent rock; (C) Back-scattered electron images of the vein-like serpentine in the saprolite and associated Ni concentrations (wt%).
(Mineral abbreviations: Srp—serpentine, Chr—chromite).

250
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Table 4
Representative EMP analysis (wt%) and structural formulae (in atoms per formula unit) of the serpentine in the saprolite samples.

Type Ni-serpentine

Sample S-1 S-1 S-1 S-2 S-2 S-2 S-3 S-3 S-4

Spot 1 2 3 4 5 6 7 8 9

Si 19.32 19.74 18.04 19.98 20.43 19.81 19.53 20.02 19.96


Ti 0.13 < 0.01 < 0.01 0.01 < 0.01 < 0.01 0.01 0.01 < 0.01
Al 0.28 0.18 0.14 0.50 1.06 0.16 0.45 0.48 0.16
Fe 7.96 7.87 6.51 5.67 6.51 3.98 5.87 5.76 4.06
Mg 14.46 13.82 18.78 19.85 15.67 20.96 19.79 19.87 20.36
Ni 3.51 4.00 2.60 1.49 3.09 1.18 1.49 1.83 1.20
Cr 0.15 0.10 0.03 0.03 0.36 0.02 0.03 0.03 0.02
Mn 0.09 0.02 0.03 0.01 0.05 0.02 0.01 0.01 0.02
Co < 0.01 < 0.01 < 0.01 < 0.01 0.03 < 0.01 < 0.01 < 0.01 < 0.01
Ca 0.21 0.26 0.11 0.06 0.19 0.06 0.06 0.05 0.11
K 0.02 0.08 0.02 < 0.01 0.03 < 0.01 < 0.01 < 0.01 < 0.01
Na 0.04 0.05 < 0.01 0.01 0.04 0.02 0.01 0.01 0.02
Si 2.18 2.23 2.01 2.07 2.17 2.07 2.05 2.07 2.08
Ti 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01
Al 0.03 0.02 0.02 0.06 0.12 0.02 0.06 0.06 0.02
ΣTetr. 2.22 2.25 2.03 2.13 2.29 2.09 2.11 2.13 2.10
Fe 0.45 0.45 0.36 0.30 0.35 0.21 0.30 0.29 0.22
Mg 1.89 1.81 2.43 2.39 1.93 2.55 2.39 2.36 2.53
Ni 0.19 0.22 0.14 0.07 0.16 0.06 0.07 0.09 0.05
ΣOct. 2.55 2.48 2.94 2.76 2.45 2.82 2.76 2.69 2.80
Ca 0.02 0.02 0.01 < 0.01 0.01 < 0.01 < 0.01 < 0.01 < 0.01
K < 0.01 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01
Na < 0.01 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01 < 0.01

the studied case is slightly different from garnierite of the Soroako laminated, spherical, or botryoidal texture. In contrast, the serpentine-
deposit in Indonesia, which also contains sepiolite-like phases besides like minerals are much bigger than those talc-like ones in grain size,
the serpentine-like and talc-like phases (Sufriadin et al., 2015). More- with a few to several hundreds of micrometers. They are sheet, needle-
over, there are several other garnierite examples which present sig- like, and tubular in single crystal shape, and their aggregates are vein-
nificant difference with this studied case. For instance, the garnierite like, mesh-like, and booklets in microtexture, similar to those ob-
from the Falcondo deposit in Dominican Republic is identified as an servations from many studied cases (e.g., Gleeson et al., 2004; Sufriadin
example of sepiolite-like phase dominant (Tauler et al., 2009; et al., 2015; Wells et al., 2009). Different from this case, some ser-
Villanova-de-Benavent et al., 2016), whereas the garnierite from the pentine-like and talc-like phases occur as intimate mixtures and they
Koniambo deposit in New Caledonia is determined as an example of cannot be separated by hand picking even under the optical micro-
simple talc-like phases (Cathelineau et al., 2016). scope, such as in the Falcondo deposit, Dominican Republic (e.g.,
Large compositional and textural variations are observed in the Villanova-de-Benavent et al., 2014).
garnierite samples from the study site. It seems that the relative pro- Chemical compositions of both serpentine-like and talc-like gar-
portions of serpentine-like and talc-like phases function as an important nierite-forming minerals show a typical isomorphous series exists be-
factor regulating the petrographic appearance of garnierite. tween ideal Mg and Ni end-member phases, which is consistent with
Empirically, if the mineralogy is dominated by the serpentine-like current understanding of Ni-Mg-bearing phyllosilicates in garnierite
phase, garnierite usually shows a typical pale-green color, loose and ores (e.g., Brand et al., 1998; Brindley, 1978; Freyssinet et al., 2005;
fragile texture in hand specimen. In contrast, if the talc-like phase is the Gleeson et al., 2003, 2004; Springer, 1974, 1976; Wells et al., 2009).
dominant mineralogy, garnierite is characterized by a strikingly bright- Microprobe chemical data indicate that the serpentine-like minerals can
green or jade-green color, compact and hardened texture. This finding be classified into the lizardite-népouite series, while the talc-like mi-
indicates that the color and texture characteristic of garnierite may nerals can be grouped as the kerolite-pimelite series. Theoretically,
provide an empirical reference to estimate the relative proportion of pure népouite and pimelite can have Ni concentrations up to 46.3 wt%
serpentine-like or talc-like phases. Also, Brindley and Hang (1973) and 35.2 wt% individually, but none of them has been found yet in
suggested that the green color of garnierite is related to its Ni con- previous and this study. Changes in Ni and Mg contents were significant
centration, and a color index could be inferred based on the Munsell both in the serpentine-like and talc-like minerals even in micro-area,
color charts. However, Villanova-de-Benavent et al. (2014) suggest that which may suggest a discontinuum of Ni for Mg solid solution between
intensity of the greenish color is not a direct indicator of the Ni con- end-member phases. The elemental mapping analyses indicate that Ni is
centration in the garnierite sample, but gives information on the preferentially enriched in talc-like phases (7.84–23.54 wt% Ni) com-
dominant mineral phase present in the garnierite. pared to serpentine-like phases (mostly ranging in 2.63–8.49 wt% Ni) in
Textural relationship of the serpentine-like and talc-like phases is studied samples, similar to the observation of Villanova-de-Benavent
quite complex. The studied samples show that these two phases occur et al. (2016). However, the mechanism of Ni enrichment in the ser-
together in the same sample and they could not be well distinguished by pentine-like and talc-like garnierite-forming minerals is not yet clear.
hand picking as indicated by the results of X-ray diffraction analysis.
However, at least part of them could be recognized and separated under
optical or electron microscope according to their distinctive miner- 5.2. Paragenetic sequence of garnierite-forming minerals
alogical characteristics. The talc-like garnierite-forming minerals are
fairly fine in their crystal size, partly falling to the nanoscale. Their Although the mineral composition of the studied garnierite appears
aggregates show a specific colloform microtexture, such as the relatively simple, the paragenetic sequence of the garnierite-forming
minerals is more complicated than it appears. Under the optical and

251
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

electron microscope, the Ni-lizardite aggregate is usually coated by talc- Mg2+ and Si4+) weathering solutions (Soler et al., 2008; Villanova-de-
like phases, while the talc-like mineral aggregate is infilled frequently Benavent et al., 2014, 2016; Wells et al., 2009). A basic question that
by microcrystalline quartz matrix. These textural relationships suggest arises is: what associations may exist between the garnierite formation
that the Ni-lizardite is formed prior to the talc-like phases and quartz. and regolith hydrology? Specifically, what kind of water flow regime
There is no debate that the talc-like phases and quartz with typical may account for the garnierite-forming weathering solution, and how
colloidal texture are of secondary precipitation origin, similar to the might such hydrologic regime be deciphered?
conclusion from many case studies (e.g., Cathelineau et al., 2016; Wells It is recognized that matrix flow and preferential flow are two main
et al., 2009). The formation mechanism of Ni-lizardite, however, ap- flow patterns in most soil environments (Beven and Germann, 1982;
pears not to be of the same origin. Mineral chemistry data show that the Lin, 2010). Matrix flow refers to water flow through soil matrix where
Ni-lizardite contain high Fe concentrations (2.03–6.44 wt%). Manceau water percolates through the pore space uniformly. It is a mostly uni-
and Calas (1985) suggested that the garnierite-forming minerals origi- versal hydrologic flow pattern linked to primary mineral weathering
nated by the secondary precipitation process commonly contain low Fe and secondary mineral formation (Lin, 2010). Preferential flow refers to
concentrations (< 0.5 wt%). The reason is that the migration ability of the process whereby water moves by preferred pathways in an ac-
Fe is weak in lateritization, prohibiting it to enter into the aqueous celerated speed through a fraction of a porous medium, thus bypassing
solution. And not only that, we notice that the secondary Ni-serpentine a portion of soil matrix (Clothier et al., 2008; Lin, 2010). In contrast to
in garnierite possess significant similarity to those inherited serpentines matrix flow, preferential flow is more closely related to the non-uni-
observed in saprolite as shown in Fig. 9. They both show typical mesh- formity and heterogeneous organization of the porous media (Lin,
like texture, tabular or vein-like shape, and no significant difference is 2010). Numerous previous studies have demonstrated that preferential
observed in their grain size. This suggests that the Ni-lizardites are not flow can occur in practically all natural soils and landscapes (e.g.,
formed by secondary precipitation but by inheritance from the sapro- Hagedorn and Bundt, 2002; Lin, 2010; Morales et al., 2010; Zhang
lite. Their high Ni-contents can be explained by an ion exchange process et al., 2016), and it can be identified by using various methods (e.g.,
between weathering solution and serpentine. Magnesium ion of ser- Graham and Lin, 2011; Guo et al., 2014; Lin and Zhou, 2008). An in-
pentine is liable to exchange with Ni ion of the leached solution due to creasing number of studies have found that preferential flow may have
their similar ionic radium (Ni = 0.08 nm, Mg = 0.077 nm) (Golightly, a deep impact on soil chemistry and biology, and so, the chemical and
1979; Gallardo et al., 2010). Therefore, a successive paragenetic se- biological characteristics of preferential flow pathways differ from
quence is suggested following as Ni-lizardite, kerolite-pimelite and those in soils governed by a matrix flow (e.g., Bundt et al., 2001;
quartz. Serpentine with elevated Fe may have been inherited from the Hagedorn and Bundt, 2002; Lipsius and Mooney, 2006). Therefore, it is
saprolite in a first instance and enriched in Ni by cation exchange possible to infer a present or past preferential flow and its pathways by
processes. Following that, kerolite-pimelite is formed by secondary decoding the information recorded in regolith. Some potential in-
precipitation process, and eventually quartz is precipitated as the dicators of preferential flow may include macropores (including frac-
terminal phase. tures and joints), chemical and mineralogical compositions of regolith
Gali et al. (2012) proposed a model to interpret the paragenetic constitutes, and even the speciation of elements in soil or regolith
sequence of garnierite-forming minerals based on the assumption that materials (e.g., Bundt et al., 2001; Bogner et al., 2012; Garrido and
in an Al-free system. The stability of garnierite-forming minerals is Helmhart, 2012; Jarvis, 2007; Lin, 2010). Here, we present evidence to
mainly controlled by the silica activity. As a result, the ideal formation support such a hypothesis in the context of garnierite occurrence; that
of the Ni ore occurs as a successive precipitation of mineral phases is, the occurrence of garnierite might be located within preferential
(serpentine series → talc series → sepiolite series → quartz) progres- flow regime.
sively enriched in Ni and Si, because Si activity increases with time and Field observations demonstrate that the occurrence of garnierite is
through the profile. Comparing with this thermodynamic model, the fairly heterogeneous. It is only developed in some fractures, joints, and
above suggested sequence of the garnierite-forming minerals at Ko- other types of macrospores in the lower part of the saprolite horizon.
londale is basically consistent with the theoretical assumption. These locations are commonly thought as the most probable pathways
Another interesting fact is that minor amount of népouite is ob- of preferential flow through regolith (Beven and Germann, 1982;
served occurring as infillings along shrinkage cracks at the internal Jarvis, 2007). The microtexture of garnierite infillings hosted in frac-
space of the kerolite-pimelite as well as at the boundary between ker- tures (Fig. 5B) suggests that parts of faults and joints may be active
olite-pimelite and quartz (Fig. 4F). This indicates that the népouite is during the weathering processes, and can be defined as syn-weathering
precipitated after talc-like phase and quartz. However, it is in contra- structures. Similar observations of the syn-weathering structure asso-
diction with the thermodynamic model. Also, quartz is identified with ciated with garnierite are also presented in New Caledonia by Cluzel
more than one stage. Besides the above referred microcrystalline quartz and Vigier (2008) as well as in Dominican Republic by Villanova-de-
occurring at a same stage with the talc-like, there are also many quartz Benavent et al. (2014). One could think that a syn-weathering structure
matrices as cement of the brecciated talc-like aggregate (Fig. 5B). To is more likely to have more open space than an inactive one, which
interpret these above facts, one possible explanation is that the né- could favor an effective hydraulic conductivity, higher infiltration and
pouite and part of quartz may represent another stage of paragenetic percolation rates, and thus become a preferential flow path for the
sequence; i.e., two stage paragenetic sequences of mineral precipitation solutions.
during the formation of garnierite is suggested. The late stage sequence Throughout the lateritic profile studied there are some ferruginous
is following as népouite and quartz, which overlapped the first stage pipes (Fig. 2A) mainly composed of Fe-oxides. They may penetrate
sequence. vertically to a deep depth, even down to the lower saprolite horizon. In
addition, numerous patches of Fe- and Mn-oxides are observed sur-
5.3. Hydrologic consideration of garnierite occurrence rounding the garnierite veins (Fig. 2C). These patches decrease pro-
gressively far from the garnierite occurrences. White et al. (2005)
Water is one of the main driving forces and agents of physical and suggested that the presence of mottling and vertical stringers of sec-
chemical weathering, and it is well recognized that water plays a par- ondary Fe and Mn oxides as well as translocated clay minerals can be
ticular important role in soil or regolith mineralogy (Lin, 2010; Scott taken as evidence of preferential flow pathways. We consider that the
and Pain, 2008; Taylor and Eggleton, 2001). However, an under- occurrences of ferruginous pipes, Fe-oxides and Mn-oxides mineral
standing of the link between garnierite formation and hydrologic pro- patches are all linked to the mobilization and downward transport of
cesses remains unclear. As a secondary weathering material, the for- metal elements during laterization. Iron and Mn eluviate downwards in
mation of garnierite is considered to be governed by cation-rich (Ni2+, oxidizing and acidic conditions mainly from the ferruginous and

252
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

limonite horizons to the lower saprolite throughout vertical fractures as solution is oversaturated at some places within preferential flow path-
preferential pathways. Moreover, similar to the observations in the ways.
Soroako garnierite-hosting profile (Golightly, 1981), we notice that Geographically, the study site within a tropical rainforest landscape
there is minor secondary quartz coexisting with garnierite. Since sec- is a favorable place for generating a high-frequent preferential flow
ondary quartz is usually formed in acidic conditions, its occurrence in a phenomena within soil/regolith (Lin, 2010), because such regions
deep regolith can be taken as mineral evidence for the rapid access of usually possess high rainfall and pronounced subsurface flowpaths.
surface acidic weathering solution to deep depths along preferential Macropore network, probably connected by combining both biological
flow pathways. channels (e.g., tree roots, worm holes, etc.) and geological channels
In contrast to saprolite samples, the garnierite shows organic matter (e.g., fractures, joints, soil cracks, etc.), may serve as a preferential
concentrations (27.9 g/kg vs 19.5 g/kg on average) but lower pH values pathway for a rapid percolation of garnierite-forming solution. Notably,
(6.62 vs 7.09 on average), and these two features can be explained as those syn-weathering structures described above may function as su-
well by the specific physicochemical environment of preferential flow perhighways, even if those types of macropores only occur in the sa-
pathways. As for organic matter, Kuzyakov and Domanski (2000) sug- prolite horizon and account for a minor portion of the total porosity
gested that rhizodeposition of organic compounds, decomposition of within the regolith. Preferential flow could trigger strong leaching and
dead roots, and transport of dissolved organic carbon from organic result in an enhanced elemental mobilization along flow pathways (e.g.,
horizons are major sources of organic C input to the soil. Thus, its in- Backnäs et al., 2012; Bogner et al., 2012; Garrido and Helmhart, 2012).
creased concentration in garnierite could be explained by enhanced The majority of the garnieite-forming elements (Ni, Mg, and Si) may be
transport of dissolved organic carbon via a preferential flow pathway. gained along preferential flow pathways rather than the whole regolith.
Bundt et al. (2001), for example, found that soil organic matter con- At the upper regolith, the Ni concentration in limonite is empirically
centration is 41% higher in preferential flow pathways than that in soil ranging around 0.5–1.5 wt%, and Ni is prone to be released from li-
matrix. Also, Bundt et al. (2001) pointed out that preferential flow monite into percolating water under an acid condition (Freyssinet et al.,
pathway is a hotspot for microbial activity, which also likely has con- 2005). Downward to the lower regolith, more Ni, Mg and Si could be
tributed to higher organic matter concentration in such hotspots. With added into percolating solution with the dissolution of serpentine
regard to lower pH values, we suppose that the percolation of acid soil (0.3–3.0 wt% Ni) in the saprolite horizon (Golightly, 1981). Carbonic
solution from organic horizons along preferential flow pathways play acid bounded or complexed with an organic acid is suggested as the
an important role. main form of Ni2+ transport (Gleeson et al., 2003; Retallack, 2010).
Overall, while we do not have direct hydrologic data on the pre- However, given the colloidal textures observed in the garnierite-
ferential flow in the studied profile, several lines of evidence (garnierite forming minerals, the form of colloid-mediated Ni transport should be
field occurrence pattern, the nature of garnierite-hosting structure, taken into consideration as well. Theoretically, Pilgrim and Huff (1983)
mineral association with garnierite, pH value and organic matter con- as well as Hardy et al. (2000) suggested that the colloidal particles are
centration) all indicate that preferential flow may have been active in detached at or near the soil surface by raindrop impact and then sub-
the studied profile, and the garnierite occurrence could be associated surface flow subsequently carries the colloidal particles primarily
with a preferential flow regime. through soil macropores. Contrastingly, since colloids are easily coa-
gulated, filtered and then are less mobile when they move in fine pores
5.4. A garnierite formation model (McCarthy and McKay, 2004), matrix flow is far less important in col-
loid transport than preferential flow. Thus, we believe that colloid-
There is no debate on the initial Ni source of garnierite from its mediated Ni transport along preferential pathways may have played a
parent rock, as it is widely observed that less-serpentinized or un- significant role in total Ni flux through the studied regolith.
serpentinized peridotites seem to be the most favorable parent rocks for Solute unloading from leached weathering solution occurs when the
the development of a garnierite-hosting regolith, as reported in New physicochemical condition of a preferential pathway is changed (Scott
Caledonia (e.g., Butt and Cluzel, 2013; Trescases, 1986; Wells et al., and Pain, 2008). The garnierite-forming minerals are formed by a
2009) and Caribbean area (e.g., Gleeson et al., 2003; Lewis et al., secondary precipitation process as leached solution reaches a locally
2006). Prendergast (2013) suggested that in some areas of the Great over-saturated state (Birsoy, 2002; Manceau and Calas, 1985). This
Dyke, Zimbabwe, the neoformation of garnierite-type minerals may process is probably pH-dependent, and it could occur where there is a
have been promoted by the high proportion of olivine in parent rocks. sudden change in pH from an acidic to an alkaline environment within
Such parent rocks may possess a high Ni background concentration the weathering profile downward from limonite horizon to saprolite
(usually ≥0.3 wt% Ni), offering a good metal source for the formation horizon (Golightly, 1979). In addition, other mechanisms for garnierite
of secondary Ni-bearing minerals through weathering. In contrast, the precipitation are also proposed by linking garnierite formation to the
parent rock in this study site is serpentinite, which does not seem to be residence time of water or water evaporation process, which could re-
an ideal rock type accounting for the garnierite formation because of its sult in an increase in cation concentration of solution and eventually
relative lower background Ni concentration (0.2 wt% Ni). Thus, under fulfill an over-saturated state (Cathelineau et al., 2016). We believe,
this specific geological background, we infer that strong Ni mobiliza- based on the colloid-mediated property of garnierite-forming minerals,
tion and redistribution processes may take place during the laterization that the colloid retention mechanism should also be considered in this
of serpentinite. Only a strong remobilization of Ni could have provided case. Wei et al. (2012) pointed out that the mechanism of colloidal film
sufficient Ni fluxes required for garnierite formation. Based on the straining and air-water interface capturing play significant roles in
above findings of spatial linkage between the garnierite occurrence and colloid retention at fractures whenever film flow occurs or gas bubbles
preferential flow regime as well as the significant impact of preferential appear under unsaturated conditions. Besides that, discontinuities in
flow on regolith development, we propose that the formation of gar- preferential flow pathways, such as dead-end macropores (Allaire-
nierite in the study site could be interpreted to relate to a preferential Leung et al., 2000), may also be another possible reason for colloid
flow-induced model (Fig. 10). The main idea here is that preferential retention and result in local Ni precipitation. Hence, we speculate that
flow may act as an essential driving force and agent of garnierite- preferential flow pathways act as both a conduit for Ni transport into
forming element (Ni, Mg and Si) mobility, which could trigger a strong deeper horizons and a sink for Ni precipitation. Moreover, due to an
element activation and transportation process along the preferential episodic and pulsed nature of preferential flow, the secondary pre-
flow pathways, in particular promoting a Ni geochemical redistribution cipitation process might be discontinuous, and as a consequence the
from dispersed state to enriched state by weathering solution, and ul- precipitation products will possess variable chemical compositions at
timately result in the precipitation of garnierite-forming minerals when different stages. This hypothesis is supported by the chemistry of the

253
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Fig. 10. Proposed model of the garnierite formation induced by preferential flow. (A) Two distinguished kinds of hydrologic flow regimes in lateritic regolith, and the preferential flow
regime linked to the percolation and precipitation of garnierite-forming weathering solution; (B) Formation mechanism of the garnierite-forming minerals as leached solution reaches a
locally over-saturated state likely caused by entering into dead-end macropores in the lower part of regolith.

talc-like garnierite-forming phases, which exhibit a fairly fluctuated 6. Conclusions


elemental distribution pattern in their oscillation laminas (Fig. 8A–B).
As a whole, we conclude from this proposed model that the (1) Garnierite from the Kolonodale area, Sulawesi Island, Indonesia is
weathering solutions accounting for the garnierite formation do not defined as an intimate mixture of Ni-rich serpentine- and talc-like
penetrate the whole regolith, but percolate only along preferential phases. The majority of the talc-like phases with colloidal micro-
pathways. In other words, the garnierite formation is genetically asso- textures can be precisely defined as kerolite, along with a minor
ciated with the preferential flow process which occurs at a limited part amount of pimelite. Most of the serpentine-like phases with mesh-
of regolith rather than the whole regolith. The Ni geochemical evolu- and booklets-like microtextures are Ni-lizardite, and a few népouite
tion induced by preferential flow could give an explanation of the non- with laths-like microtexture is identified as well. Variation of che-
uniform pattern of the garnierite field occurrence. In addition, from a mical composition in garnierite-forming minerals covers a large
hydrologic perspective, the contributions of matrix flow in secondary interval mainly between Mg and Ni, which may reflect the episodic
mineral formation are undeniable in any weathering profiles (Lin, changes of solution chemistry.
2010). In this studied regolith, except the garnierite-forming minerals, (2) Textural evidences indicate the paragenetic sequence of the gar-
some other Ni-bearing secondary minerals with relative uniform pat- nierite-forming minerals involves two stages. The first stage is Ni-
tern (i.e., Ni-serpentine in saprolite and Ni-goethite in limonite) are lizardite, kerolite-pimelite and quartz. Serpentine with elevated Fe
assumed to be linked with matrix flow, which is more universal within may have been inherited from the saprolite in a first instance and
regolith and accounts for the majority of secondary minerals present in enriched in Ni by cation exchange processes. Newly formed mi-
the whole profile (Fig. 10). nerals by secondary precipitation process are kerolite-pimelite and
In a broad sense, this study offers further insights into the link be- then followed by microcrystalline Ni-free quartz, which coated the
tween preferential flow and secondary mineral formation. Many issues previously formed lizardite aggregates. In addition, based on the
regarding the secondary minerals of regolith, especially for those with specific occurrence of népouite as infillings along shrinkage cracks
heterogeneous distribution, could be explained by using the theory of within previously formed minerals, we assume that a second stage
preferential flow. Such a recognition has been preliminarily alluded to is népouite and quartz, but overlapping the first stage.
by several studies (e.g., Garrido and Helmhart, 2012; Taylor and (3) Several lines of evidence indicate that the occurrence of garnierite
Eggleton, 2001), but has lacked proof. However, further work is needed could be linked to a preferential flow hydrologic regime. Main in-
to bridge long-term secondary mineral formation and short-term hy- dicators include: (i) non-uniform pattern of the garnierite field oc-
drologic processes by taking into account more theoretical and case currence, in accordance with the style of preferential flow activity;
studies. (ii) syn-weathering active nature of the garnierite-hosting struc-
tures, serving as accessible conduits of preferential flow movement;
(iii) close relationship between the garnierite occurrence and ver-
tical FeeMn oxides pipes as well as FeeMn oxides patched areas,

254
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

reflecting strong metal leaching from the upper horizons induced by van der Ent, A., Baker, A.J.M., van Balgooy, M.M.J., Tjoa, A., 2013. Ultramafic nickel
preferential flow; (iv) specific physico-chemical property of the laterites in Indonesia (Sulawesi, Halmahera): mining, nickel hyperaccumulators and
opportunities for phytomining. J. Geochem. Explor. 128, 72–79.
garnierite location with higher organic matter concentrations but Faust, G.T., 1966. The hydrous magnesium silicates-the garnierite group. Am. Mineral.
lower pH values, similar to the common feathers of soil/regolith 51, 279–298.
domains affected by preferential flow. Freyssinet, P., Butt, C.R.M., Morris, R.C., Piantone, P., 2005. Ore-forming processes re-
lated to lateritic weathering. In: Economic Geology 100th Anniversary Volumepp.
(4) A preferential flow induced model is proposed to explain the origin 681–722.
of garnierite by coupling an element geochemical process and a Fritsch, E., Juillot, F., Dublet, G., Fonteneau, L., Fandeur, D., Martin, E., Caner, L.,
hydrological process at the profile scale. Preferential flow is con- Auzende, A.L., Grauby, O., Beaufort, D., 2016. An alternative model for the formation
of hydrous Mg/Ni layer silicates (‘deweylite’/‘garnierite’) in faulted peridotites of
sidered to be actively involved in Ni mobilization and redistribution New Caledonia: I. Texture and mineralogy of a paragenetic succession of silicate
during serpentinite lateritization. It serves as a main driving force infillings. Eur. J. Mineral. 28 (2), 295–311.
and agent of element mobility. Fractures identified as being of a Fu, W., Yang, J.W., Yang, M.L., Pang, B.C., Liu, X.J., Niu, H.J., Huang, X.R., 2014.
Mineralogical and geochemical characteristics of a serpentinite-derived lateritic
syn-weathering active nature may act as significant preferential
profile from East Sulawesi, Indonesia: implications for the lateritization process and
pathway for a rapid percolation of garnerite-forming solution. Ni supergene enrichment in the tropical rainforest. J. Asian Earth Sci. 93, 74–88.
Along the preferential pathway, enhanced leaching process may Gali, S., Soler, J.M., Proenza, J.A., Lewis, J.F., Cama, J., Tauler, E., 2012. Ni enrichment
provide a sufficient element flux required for the garnierite for- and stability of Al-free garnierite solid-solutions: a thermodynamic approach. Clays
Clay Miner. 60, 121–135.
mation from a serpentinite with low Ni concentration (< 0.2 wt% Gallardo, T., Tauler, E., Proenza, J.A., Lewis, J.F., Gali, S., Labrador, M., Longo, F., Bloise,
Ni). G., 2010. Geology, mineralogy and geochemistry of the Loma Ortega Ni lateritic
deposit, Dominican Republic. Macla 13, 89–90.
Garrido, F., Helmhart, M., 2012. Lead and soil properties distributions in a roadside soil:
Acknowledgments effect of preferential flow paths. Geoderma 170, 305–313.
Gleeson, S.A., Butt, C.R.M., Elias, M., 2003. Nickle lateritics: a review. Newsl. Soc. Econ.
Research was financially supported by the National Natural Science Geol. 4, 12–18.
Gleeson, S.A., Herrington, R.J., Durango, J., Velázquez, C.A., 2004. The mineralogy and
Foundation of China (41462005; 41102051), Guangxi Natural Science geochemistry of the Cerro Matoso S.A. Ni lateritic deposit, Montelíbano, Colombia.
Foundation (2014GXNSFAA118304), and the project of Collaborative Econ. Geol. 99, 1197–1213.
Innovation Center for Exploration of Hidden Nonferrous Metal Deposits Golightly, J.P., 1979. Geology of Soroako nickeliferous lateritic deposits. In: Evans, D.J.I.,
Shoemaker, R.S., Veltman, H. (Eds.), International Lateritic Symposium: New York.
and Development of New Materials in Guangxi. We acknowledge the Artificial Intelligence in Medicine, pp. 38–56.
support of the PT. Pan China Indon Co., Ltd. for our field investigations. Golightly, J.P., 1981. Nickeliferous lateritic deposits. In: Economic Geology, 75th
We also would like to thank Prof. Bryan Krapez from Curtin University, Anniversary Volumepp. 710–735.
Golightly, J.P., 2010. Progress in understanding the evolution of nickel lateritics. In:
Prof. Li Jianwei from China University of Geoscience, and Prof. Feng
Goldfarb, R.J., Marsh, E.E., Monecke, T. (Eds.), The Challenge of Finding New
Zhuohai from Guilin University of Technology for their assistance. This Mineral Resources-Global Metallogeny, Innovative Exploration, and New Discoveries.
manuscript has benefited from discussions with Dr. Yunqiang Wang of Society of Economic Geologists Special Publication Vol. 15. pp. 451–485.
the Chinese Academy of Sciences and Dr. Guo Li of Pennsylvania State Graham, C.B., Lin, H.S., 2011. Controls and frequency of preferential flow occurrence: a
175-event analysis. Vadose Zone J. 10, 816–831.
University. We appreciate careful and exhaustive comments and revi- Guo, L., Chen, J., Lin, H., 2014. Subsurface lateral preferential flow network revealed by
sions of three anonymous journal reviewers, which improved the time-lapse ground-penetrating radar in a hillslope. Water Resour. Res. 50,
manuscript. 9127–9147.
Hagedorn, F., Bundt, M., 2002. The age of preferential flow paths. Geoderma 108,
119–132.
References Hall, R., Wilson, E.J., 2000. Neogene suture in Eastern Indonesia. J. Asian Earth Sci. 18,
781–808.
Hardy, I.A.J., Carter, A.D., Leeds-Harrison, P.B., Foster, I.D.L., Sanders, R.M., 2000. The
Allaire-Leung, S.E., Gupta, S.C., Moncrief, J.F., 2000. Water and solute movement in soil
origin of sediment in field drainage water. In: Foster, I.D.L. (Ed.), Tracers in
as influenced by macropore characteristics: 1. Macropore continuity. J. Contam.
Geomorphology. John Wiley & Sons Ltd Chichester, New York, pp. 241–257.
Hydrol. 41 (3–4), 283–301.
Ilyasa, A., Kashiwaya, K., Koike, K., 2016. Ni grade distribution in lateritic characterized
Backnäs, S., Laine-Kaulio, H., Kløve, B., 2012. Phosphorus forms and related soil chem-
from geostatistics, topography and the paleo-groundwater system in Sorowako,
istry in preferential flowpaths and the soil matrix of a forested podzolic till soil
Indonesia. J. Geochem. Explor. 165, 174–188.
profile. Geoderma 189-190 (6), 50–64.
Jarvis, N.J., 2007. A review of non-equilibrium water flow and solute transport in soil
Berger, V.I., Singer, D.A., Bliss, J.D., 2011. Ni-Co laterite deposits of the world: database
macropores: principles, controlling factors and consequences for water quality. Eur.
and grade and tonnage models. In: U.S. Geological Survey Open-File Report. 1058.
J. Soil Sci. 58, 523–546.
Beven, K., Germann, P., 1982. Macropores and water flow in soils. Water Resour. Res. 18,
Kadarusmana, A., Miyashitab, S., Maruyamaa, S., Parkinsonc, C.D., Ishikawad, A., 2004.
1311–1325.
Petrology, geochemistry and paleogeographic reconstruction of the East Sulawesi
Birsoy, R., 2002. Formation of sepiolite-palygorskite and related minerals from solution.
Ophiolite, Indonesia. Tectonophysics 392, 55–83.
Clays Clay Miner. 50 (6), 736–745.
Kuck, P.H., 2013. Mineral resource of the month: nickel. Geotimes 51 (10), 47.
Bogner, C., Borken, W., Huwe, B., 2012. Impact of preferential flow on soil chemistry of a
Kuzyakov, Y., Domanski, G., 2000. Carbon input by plants into the soil. Review. J. Plant
podzol. Geoderma 175-176 (3), 37–46.
Nutr. Soil Sci. 163, 421–431.
Brand, N.W., Butt, C.R.M., Elias, M., 1998. Nickle lateritics: classifications and features.
Lewis, J.F., Draper, G., Proenza, J.A., Espaillat, J., Jimenez, J., 2006. Ophiolite-related
AGSO J. Aust. Geol. Geophys. 17 (4), 81–88.
ultramafic rocks (serpentinites) in the Caribbean region: a review of their occurrence,
Brindley, G.W., 1978. The structure and chemistry of hydrous nickel containing silicate
composition, origin, emplacement and Ni-lateritic soils formation. Acta Geol. Hung.
and aluminate minerals. Bull. Bur. Rech. Géol. Min. II 3, 233–245.
4, 237–263.
Brindley, G.W., Hang, P.T., 1973. The nature of Garnierites—I: structures, chemical
Lin, H., 2010. Linking principles of soil formation and flow regimes. J. Hydrol. 393, 3–19.
compositions and color characteristic. Clays Clay Miner. 21, 27–40.
Lin, H.S., Zhou, X., 2008. Evidence of subsurface preferential flow using soil hydrologic
Brindley, G.W., Bish, D.L., Wan, H.M., 1977. The nature of kerolite, its relation to talc and
monitoring in the Shale Hills Catchment. Eur. J. Soil Sci. 59, 34–49.
stevensite. Mineral. Mag. 41, 443–452.
Lipsius, K., Mooney, S.J., 2006. Using image analysis of tracer staining to examine the infi
Brindley, G.W., Bish, D.L., Wan, H.M., 1979. Composition, structures, and properties of
ltration patterns in a water repellent contaminated sandy soil. Geoderma 136,
nickel-containing minerals in the kerolite-pimelite series. Am. Mineral. 64, 615–625.
865–875.
Bundt, M., Widmer, F., Pesaro, M., Zeyer, J., Blaser, P., 2001. Preferential flow paths:
Manceau, A., Calas, G., 1985. Heterogeneous distribution of nickel in hydrous silicaes
biological “hot spots” in soils. Soil Biol. Biochem. 33, 729–738.
from New Caledonia ore deposits. Am. Mineral. 70 (5), 549–558.
Butt, C.R.M., Cluzel, D., 2013. Nickel lateritic ore deposits: weathered serpentinites.
McCarthy, J.F., McKay, L.D., 2004. Colloid transport in the subsurface: past, present, and
Elements 9, 123–128.
future challenges. Vadose Zone J. 3, 326–337.
Cathelineau, M., Quesnel, B., Gautier, P., Boulvais, P., Couteau, C., Drouillet, M., 2016.
Morales, V.L., Parlange, J.-Y., Steenhuis, T.S., 2010. Are preferential flow paths perpe-
Nickel dispersion and enrichment at the bottom of the regolith: formation of pimelite
tuated by microbial activity in the soil matrix? A review. J. Hydrol. 393, 29–36.
target-like ores in rock block joints (Koniambo Ni deposit, New Caledonia). Mineral.
Mubroto, B., Briden, J.C., Mcclelland, E., Hall, R., 1994. Palaeomagnetism of the Balantak
Deposita 51, 271–282.
ophiolite, Sulawesi. Earth Planet. Sci. Lett. 125 (1–4), 193–209.
Clothier, B.E., Green, S.R., Deurer, M., 2008. Preferential flow and transport in soil:
Mudd, G.M., 2010. Global trends and environmental issues in nickel mining: sulfides
progress and prognosis. Eur. J. Soil Sci. 59, 2–13.
versus lateritic. Ore Geol. Rev. 38, 9–26.
Cluzel, D., Vigier, B., 2008. Syntectonic mobility of supergene nickel ores of New
Nickel, E.H., 1992. Solid solutions in mineral nomenclature. Miner. Petrol. 46 (1), 49–53.
Caledonia (Southwest Pacific): evidence from faulted regolith and garnierite veins.
Pilgrim, D.H., Huff, D.D., 1983. Suspended sediment in rapid subsurface stormflow on a
Resour. Geol. 25, 112–117.
large field plot. Earth Surf. Process. Landf. 8, 451–463.

255
W. Fu et al. Journal of Geochemical Exploration 188 (2018) 240–256

Prendergast, M.D., 2013. Landscape evolution, regolith formation and nickel lateritic (3), 397–398.
development in the northern part of the Great Dyke, Zimbabwe. S. Afr. J. Geol. 116 Trescases, J.J., 1986. Nickeliferous lateritics: a review of the contributions of the last
(2), 219–240. 10 years. Geol. Surv. India Mem. 120, 51–62.
Retallack, G.J., 2010. Lateritization and bauxitization events. J. Econ. Geogr. 105, Varlakov, A.S., 1992. Contact metamorphism of ultramafics of the Ufalei massif, the
655–667. Southern Urals. Izv. Akad. Nauk Ser. Geol. 5, 27–38.
Scott, K.M., Pain, C.F., 2008. Regolith Science. CSIRO Pub, Dordrecht, Netherlands, Villanova-de-Benavent, C., Proenza, J.A., Galí, S., Garcia-Casco, A., Tauler, E., Lewis, J.F.,
Collingwood, Australia. Longo, F., 2014. Garnierites and garnierites: textures, mineralogy and geochemistry
Soler, J.M., Cama, J., Gali, S., Meléndezc, W., Ramírezc, A., Estangad, J., 2008. of garnierites in the Falcondo Ni-lateritic deposit, Dominican Republic. Ore Geol.
Composition and dissolution kinetics of garnierite from the Loma de Hierro Ni-la- Rev. 58, 91–109.
teritic deposit, Venezuela. Chem. Geol. 249, 191–202. Villanova-de-Benavent, C., Nieto, F., Viti, C., Proenza, J.A., Galí, S., Roqué-Rosell, J.,
Springer, G., 1974. Compositional and structural variation in garnierites. Can. Mineral. 2016. Ni-phyllosilicates (garnierites) from the Falcondo Ni-lateritic deposit
12, 381–388. (Dominican Republic): mineralogy, nanotextures, and formation mechanisms by
Springer, G., 1976. Falcondoite, nickel analogue of sepiolite. Can. Mineral. 14, 407–409. HRTEM and AEM. Am. Mineral. 101, 1460–1473.
Sufriadin, S., Idrus, A., Pramumijoyo, S., Warmada, I.W., Imai, A., 2015. Study on mi- Wei, Z., Tang, X.Y., Noam, W., Zhuo, G., 2012. A review of colloid transport in fractured
neralogy and chemistry of the saprolitic nickel ores from Soroako, Sulawesi, rocks. J. Mt. Sci. 9, 770–787.
Indonesia: implication for the lateritic ore processing. Genes Genet. Syst. 82 (6), Wells, M.A., Ramanaidou, E.R., Verrall, M., Tessarolo, C., 2009. Mineralogy and crystal
455–464. chemistry of “garnierites” in the Goro lateritic nickel deposit, New Caledonia. Eur. J.
Talovina, I.V., Lazarenkov, V.G., Ryzhkova, S.O., Ugol'kov, V.L., Vorontsova, N.I., 2008. Mineral. 21, 467–483.
Garnierite in nickel deposits of the Urals. Lithol. Miner. Resour. 43, 588–595. White, A.F., Schulz, M.S., Vivit, D.V., Blum, A.E., Stonestrom, D.A., Harden, J.W., 2005.
Tauler, E., Proenza, J.A., Gali, S., Lewis, J.F., Labrador, M., Garcia-Romero, E., Suarez, Chemical weathering rates of a soil chronosequence on granitic alluvium: III.
M., Longo, F., Bloise, G., 2009. Ni-sepiolite-falcondoite in garnierite mineralization Hydrochemical evolution and contemporary solute fluxes and rates. Geochim.
from the Falcondo Ni-lateritic deposit, Dominican Republic. Clay Miner. 44, Cosmochim. Acta 69 (8), 1975–1996.
435–454. Zhang, Y.H., Zhang, M.X., Niu, J.Z., Zheng, H.J., 2016. The preferential flow of soil: a
Taylor, G., Eggleton, R.A., 2001. Regolith geology and geomorphology. Earth-Sci. Rev. 58 widespread phenomenon in pedological perspectives. Eurasian Soil Sci. 49, 661–672.

256

You might also like