181
Genèse des concentra-ons de Nickel en Nouvelle Calédonie Compila(on des ar(cles scien(fiques publiés par GeoRessources (Labex Ressources 21)Nancy et Geosciences Rennes en collabora(on Marc Ulrich (EOST, Strasbourg) Nickel concentra-on genesis in New Caledonia Compila(on of scien(fic ar(cles published by GeoRessources (Labex Ressources 21)Nancy and Geosciences Rennes in collabora(on with Dr. Marc Ulrich (EOST, Strasbourg)

Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genèse  des  concentra-ons    de  Nickel  en  Nouvelle  Calédonie  

   Compila(on  des  ar(cles  scien(fiques  publiés  par      GeoRessources  (Labex  Ressources  21)-­‐Nancy  

   et  Geosciences  Rennes    en  collabora(on  Marc  Ulrich  (EOST,  Strasbourg)  

 

 Nickel  concentra-on  genesis  in  New  Caledonia  

   Compila(on  of  scien(fic  ar(cles  published  by    GeoRessources  (Labex  Ressources  21)-­‐Nancy  

   and  Geosciences  Rennes    in  collabora(on  with  Dr.  Marc  Ulrich  (EOST,  Strasbourg)  

       

Page 2: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 3: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genèse des concentrations de Nickel en Nouvelle Calédonie

Compilation des articles scientifiques publiés par

GeoRessources (Labex Ressources 21)-Nancy

et Geosciences Rennes

en collaboration Marc Ulrich (EOST, Strasbourg)

-----------------------------------------------------------------------------------

Nickel concentration genesis in New Caledonia

Compilation of scientific articles published by

GeoRessources (Labex Ressources 21)-Nancy

and Geosciences Rennes

in collaboration with Dr. Marc Ulrich (EOST, Strasbourg)

Page 4: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Ce volume d’articles scientifiques compile les différentes contributions des chercheurs de GeoRessources (Nancy) et Geosciences Rennes, en collaboration

avec M. Ulrich (EOST Strasbourg) dans le domaine de la genèse des concentrations de Nickel de Nouvelle Calédonie.

Ces travaux sont le fruit d’une étroite collaboration entre les chercheurs de Nancy et Rennes, initiée grâce au support logistique, technique et financier de Koniambo S.A. dans le cadre de la thèse de B. Quesnel, et de la contribution du Labex Ressources 21 dans le cadre de son programme tri-annuel sur le cycle du

Nickel

Cette compilation est réalisée à l’occasion de la tenue du Conseil scientifique du Labex Ressources 21 à Nancy en Janvier 2017.

This volume compiles the various contributions from GeoRessources (Nancy) and Geosciences Rennes researchers, in association with M. Ulrich (EOST Strasbourg) in the field of nickel concentration genesis in New Caledonia

These publications are the result of a close collaboration between the researchers of Nancy and Rennes, brought together thanks to the logistic,

technical and financial support of Koniambo S.A. within the framework of B. Quesnel PhD thesis, and through the Labex Ressources 21 annual program on

Nickel cycle

This compilation is presented for at the scientific Council of Labex Ressources 21 to be held in January 2017, in Nancy, France

Page 5: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Geology and deformation of the ophiolitic nappe in New Caledonia

1) Gautier, P., B. Quesnel, P. Boulvais, and M. Cathelineau (2016), The emplacement of the Peridotite Nappe of New Caledonia and its bearing on the tectonics of obduction, Tectonics, 35, doi:10.1002/2016TC004318..

2) Quesnel, Benoit, Gautier, Pierre, Cathelineau, Michel Boulvais Philippe, Couteau Clément, et Drouillet Maxime. (2016) The internal deformation of the Peridotite Nappe of New Caledonia: A structural study of serpentine-bearing faults and shear zones in the Koniambo Massif Journ. Struct. Geol. : 85, 51-67 .

Conditions of Magnesite formation in ophiolites

3) Quesnel, B., Gautier, P., Boulvais, P., Cathelineau, M., Maurizot, P.., Cluzel, D., Ulrich, M., Guillot, S., Lesimple, S., Couteau C. (2013) Syn-tectonic, meteoric water-derived carbonation of the New Caledonia peridotite nappe, Geology 41, 10 (2013) 1063-1066

4) Ulrich M., Munoz, M. Guillot, S., Cathelineau, M., Picard, C., Quesnel, B.Boulvais , P., et Couteau C. (2014) Dissolution–precipitation processes governing the carbonation and silicification of the serpentinite sole of the New CaledoniaOphiolite. Contrib Mineral Petrol (2014) 167:952 DOI 10.1007/s00410-013-0952-8

5) Quesnel, B., Boulvais, P., Gautier, P., Cathelineau, M., John C. M., Dierick M., Agrinier P., Drouillet M. (2016) Paired stable isotopes (O, C) and clumped isotope thermometry of magnesite and silica veins in the New Caledonia Peridotite Nappe. Geoch. Cosmochim. Acta,: 183 : 234-249

6) Quesnel, B., Boulvais, P., Gautier, P., Cathelineau, M., John C. M., Dierick M., Agrinier P., Drouillet M. (2016) Formation of silica and magnesite veins in the New Caledonian Peridotite Nappe: geometric and stable isotopes data. SGA Proceedings

Genesis and mineralogy of Ni ores

7) Cathelineau, M., Myagkiy, A., Quesnel, B., Boiron, M.-C., Gautier, P., Boulvais, P., Ulrich, M., Truche, L., Golfier, F., Drouillet, M., 2016a. Multistage crack seal vein and hydrothermal Ni enrichment in serpentinized ultramafic rocks (Koniambo massif, New Caledonia). Mineralium Deposita. doi:10.1007/s00126-016-0695-3 8) Myagkiy A., Cathelineau M., Golfier F., Truche L., Quesnel B., Drouillet, M., 2016b. Defining mechanisms of brecciation in Ni-silicate bearing fractures (Koniambo, New Caledonia). SGA Conference Nancy 2015 Proceedings, vol. 3, pp . 1181-1185.

9) Cathelineau M., Caumon, M-C, Massei F., Brie D. and Harlaux M. (2015) Raman spectra of Ni–Mg kerolite: effect of Ni–Mg substitution on O–H stretching vibrations. J. Raman ; Spectroscopy, DOI 10.1002/jrs.4746

10) Cathelineau, M., Quesnel, B., Gautier, P., Boulvais, P., Couteau, C., Drouillet, M., 2016b. Nickel dispersion and enrichment at the bottom of the regolith: formation of pimelite target-like ores in rock block joints (Koniambo Ni deposit, New Caledonia). Mineralium Deposita 51, 271–282. doi:10.1007/s00126-015-0607-

Page 6: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

11) Quesnel, B, Le Carlier de Veslud, C., Boulvais, P, Gautier, P, Cathelineau, M, Drouillet, M (2017) 3D modeling of the laterites on top of the Koniambo Massif, New Caledonia: refinement of the per descensum lateritic model for nickel mineralization. Miner Deposita, DOI 10.1007/s00126-017-0712-1 .

12) Myagkiy A., Truche, T. , Cathelineau M., Golfier F. (2017) Revealing the conditions of Ni mineralization in the laterite profiles of New Caledonia: Insights from reactive geochemical transport modelling. Chem. Geol., 466, 274-284.

Page 7: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Geology and deformation of the ophiolitic

nappe in New Caledonia

Gautier, P., B. Quesnel, P. Boulvais, and M. Cathelineau (2016)

The emplacement of the Peridotite Nappe of New Caledonia and its bearing on the tectonics of obduction,

Tectonics, 35, doi:10.1002/2016TC004318

- 1 -

Page 8: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

The emplacement of the Peridotite Nappeof New Caledonia and its bearingon the tectonics of obductionPierre Gautier1 , Benoît Quesnel1,2 , Philippe Boulvais1, and Michel Cathelineau2

1Université Rennes 1, CNRS, Géosciences Rennes UMR 6118, OSUR, Rennes, France, 2Université de Lorraine, CNRS, CREGU,GeoRessources lab, Vandoeuvre-lès-Nancy, France

Abstract The Peridotite Nappe of New Caledonia is one of the few ophiolites worldwide that escapedcollisional orogeny after obduction. Here we describe the deformation associated with serpentinizationin two klippes of the nappe in northwestern New Caledonia. The klippes are flat lying and involve S/SWvergent reverse-slip shear zones which are true compressional structures in origin. Further northeast, thenappe is folded in association with the development of a steep schistosity in low-grade metasediments.This difference in structural style indicates that the Peridotite Nappe experienced compression atgreater depths toward its root zone, suggesting a “push from the rear” mechanism of emplacement.This supports the view that the nappe has been emplaced through horizontal contraction sustained byplate convergence. We establish a crustal-scale cross section at the end of the obduction event, beforeNeogene extension. This involves a large fold nappe of high-pressure rocks bounded from below by amajor thrust. Furthermore, we show that obduction in New Caledonia occurred through dextral obliqueconvergence. Oblique convergence probably resulted from the initial obliquity between the subductiontrench and the continental ribbon that became incorporated in it. This obliquity can solve the paradoxof the Peridotite Nappe seemingly being emplaced at the same time the high-pressure rocks wereexhumed. Oblique convergence together with focused erosional denudation on the northeastern flankof the island led to exhumation of the metamorphic rocks in a steep fold nappe rising through the rear partof the orogen.

1. Introduction

Ophiolites commonly occur within orogenic belts as a result of the obduction of a piece of oceanic litho-sphere upon a continental basement [Coleman, 1971; Dewey and Bird, 1971]. Ophiolitic nappes are oftenstrongly deformed during subsequent collision tectonics and so are often not suitable for studying theinitial obduction process. The Peridotite Nappe of New Caledonia is known as an emblematic example ofan ophiolite because, as in Oman, the region has not suffered collision since the mid-Cenozoic obduction[e.g., Cluzel et al., 2012a]. However, to date, there is no consensus on the mechanism that led to the empla-cement of the Peridotite Nappe or on the tectonic evolution of New Caledonia at the time of obduction[Cluzel et al., 1995, 2001, 2012a; Rawling and Lister, 1999; Spandler et al., 2005; Baldwin et al., 2007;Lagabrielle et al., 2013]. The different models proposed so far rely essentially on data from the rock unitsunderlying the Peridotite Nappe but not on structural data from the nappe itself. Structural analysisof the peridotites has focused on high-temperature fabrics related to intraoceanic stages of deformation[Prinzhofer et al., 1980; Nicolas, 1989; Titus et al., 2011]. Only a few studies have described lowertemperature-related structures that could reflect the emplacement of the nappe or/and later tectonicevents [Guillon and Routhier, 1971; Guillon, 1975; Leguéré, 1976; Poutchovsky and Récy, 1982; Lagabrielleand Chauvet, 2008].

Recently, the deformation associated with serpentinization and carbonation of the peridotites has beencharacterized in the Koniambo Massif, one of the klippes of the Peridotite Nappe in northern NewCaledonia [Quesnel et al., 2013, 2016a]. Here we summarize the results of this analysis and present additionalobservations made in other klippes. Together with constraints from the literature about the rock unitsbeneath the Peridotite Nappe and the offshore domain, these new data yield a picture of the distributionof low-temperature deformation within the nappe that enables us to describe its mechanism of emplace-ment and the tectonics of obduction.

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 1

PUBLICATIONSTectonics

RESEARCH ARTICLE10.1002/2016TC004318

Key Points:• Compressional structures characterizethe internal deformation of thePeridotite Nappe of New Caledonia

• The nappe has been emplaced in acontext of crustal-scale contractiondriven by oblique convergence

• Obduction involved synconvergenceexhumation of HP-LT rocks in a foldnappe rising through the rear part ofthe orogen

Correspondence to:P. Gautier,[email protected]

Citation:Gautier, P., B. Quesnel, P. Boulvais,and M. Cathelineau (2016), Theemplacement of the Peridotite Nappeof New Caledonia and its bearing on thetectonics of obduction, Tectonics, 35,doi:10.1002/2016TC004318.

Received 25 JUL 2016Accepted 24 NOV 2016Accepted article online 5 DEC 2016

©2016. American Geophysical Union.All Rights Reserved.

Page 9: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

2. Geological Setting2.1. New Caledonia

New Caledonia is located in the southwest Pacific Ocean, 1300 km east of Australia (Figure 1). On the mainisland, known as the “Grande Terre,” the Peridotite Nappe overlies, with a subhorizontal tectonic contact[Avias, 1967; Guillon, 1975], a substratum composed of four main rock assemblages [Paris, 1981; Cluzelet al., 2001, 2012a].

The “basement,” forming the central part of the island consists of Permian to Early Cretaceous volcano-sedimentary rocks with a variable pre-Coniacian metamorphic overprint. This basement forms the backboneof a ≤ 100 km wide ribbon of continental crust that extends, south of the Grande Terre, into the shallowsubmarine N-S trending Norfolk Ridge.

The second group of rocks consists of Late Cretaceous (Coniacian) to late Eocene sediments. The LateCretaceous deposits unconformably overlie the basement and are interpreted as a synrift (with associated

Figure 1. Simplified structural map of the Grande Terre, New Caledonia. The distribution of ultramafic rocks is slightly modified from Maurizot and Vendé-Leclerc[2009]. Most of these rocks are originally part of the Peridotite Nappe.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 2

Page 10: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

volcanic rocks) to postrift sequence. Following Paleocene pelagic sedimentation, turbidites started toaccumulate in the north of the island during the late early Eocene at ~50Ma [Maurizot, 2011]. This changein sedimentation is interpreted as a first record of deformation in the foreland of the convergent plateboundary that eventually led to the obduction of the Peridotite Nappe. The onset of clastic sedimentationis progressively delayed further southeast [Cluzel et al., 2012a;Maurizot and Cluzel, 2014], occurring in the lateEocene near Noumea [Cluzel et al., 2001].

A third rock assemblage is represented by the Poya Terrane, with dolerites, basalts, and minor bathyalsediments representing the remnants of a Late Cretaceous to earliest Eocene oceanic floor [Cluzel et al.,2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between theoverlying Peridotite Nappe and the underlying basement and its Late Cretaceous-Eocene cover.

Finally, a fourth group of rocks is characterized by the record of a high-pressure-low-temperature (HP-LT)metamorphic overprint of Cenozoic age. These rocks are exposed in the northwestern half of the GrandeTerre (Figure 1). Within this area, a northeastward increase in metamorphic grade is recognized [e.g.,Brothers, 1974; Vitale Brovarone and Agard, 2013]. To the southwest, the first noticeable planes of schistos-ity dip at ~50° to the northeast [Maurizot et al., 1989]. Moving northeastward, the main schistosity progres-sively steepens, becoming vertical about 2.5 km before the lawsonite-in isograd, and then dips to thesouthwest [Lillie, 1970; Maurizot et al., 1989] (Figure 1). As a result, schistosities define an upright fan-likestructure with a width of about 10 km [Cluzel et al., 1995, 2001; Vitale Brovarone and Agard, 2013]. Furthernortheast, despite local complications due to some fault zones, a major structural culmination is recog-nized adjacent to the northeastern coast [Lillie, 1975; Cluzel et al., 1995, 2012a; Rawling and Lister, 1999;Vitale Brovarone and Agard, 2013] (Figure 1). Southwest of this culmination, schistosities define a largeasymmetric fold with a steeply plunging axis, the Ouégoa-Bondé fold [Lillie, 1970; Paris, 1981; Maurizotet al., 1989]. Its asymmetry is consistent with dextral shearing along a trend parallel to the culmination(Figure 1). Equivalent folds exist further northwest and southeast, slightly west of the lawsonite isograd[Paris, 1981].

Within the domain of Cenozoic metamorphism, two units are distinguished [Cluzel et al., 1995; Clarke et al.,1997]. The Pouebo Terrane, along the northern coast, is a mélange carrying blocks of mafic eclogites andminor metasediments within a matrix of talcschists. Geochemical and protolith age data indicate thatthe blocks in the mélange and the Poya Terrane are probably derived from the same oceanic domain[Cluzel et al., 2001; Spandler et al., 2005]. The Diahot Terrane, across which much of the northeastwardincrease in metamorphic grade occurs, has Late Cretaceous-Paleocene sedimentary and volcanic protolithsreminiscent of the series covering the basement [Paris, 1981; Cluzel et al., 2012a]. U-Pb ages of ~44Maon metamorphic rims of zircon grains have been interpreted as dating peak conditions in the PoueboTerrane [Spandler et al., 2005]. Other time constraints mostly consist of 39Ar-40Ar ages on white micas,of ~36.5–40Ma in the Pouebo Terrane and ~35–38Ma in the higher-grade part of the Diahot Terrane,interpreted as dating cooling of the metamorphic rocks down to ~350–450°C [Ghent et al., 1994;Baldwin et al., 2007] Using a pressure-temperature path established by Fitzherbert et al. [2004]; Baldwinet al. [2007] concluded that the above 39Ar-40Ar ages date a stage of the decompression path at around5–8 kbar. A later stage of the path is constrained by one fission track age on apatite at 34� 4Ma [Baldwinet al., 2007].

2.2. The Peridotite Nappe

The Peridotite Nappe is mainly exposed in the “Massif du Sud,” covering much of the southeastern third ofthe Grande Terre and in a series of klippes along the northwestern coast (Figure 1). In the Massif du Sud,the thickness of the nappe is at least 1.5 km but locally may reach 3.5 km [Guillon, 1975]. The nappe is mostlycomposed of harzburgites except in the northernmost klippes (Tiebaghi, Poum, and Yandé; see Figure 1 forlocation) where lherzolites are dominant [e.g., Nicolas, 1989]. In the harzburgites, compositional layering isrepresented by 1 to 100m thick layers of dunite [e.g., Guillon, 1975; Ulrich et al., 2010]. In high-elevation areasof the Massif du Sud, the harzburgites are capped by massive dunites overlain by layered wehrlites, pyroxe-nites, and gabbros, representing the base of an oceanic crust [Guillon, 1975; Prinzhofer et al., 1980; Paris, 1981].Earlier authors considered this crust as being formed at a mid-oceanic spreading ridge, but recent work hasshown that it is mainly a fore-arc crust formed in a suprasubduction zone setting [Marchesi et al., 2009; Pirardet al., 2013]. The harzburgites represent residues produced by either a single partial melting event associated

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 3

Page 11: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

with the formation of the fore-arc crust [Marchesi et al., 2009] or, more probably, at least two partial meltingevents, the first of which likely relates to a spreading ridge [Ulrich et al., 2010; Pirard et al., 2013].

In the main part of the Peridotite Nappe, the degree of serpentinization of the peridotites is moderate butvariable [Orloff, 1968], serpentines occurring preferentially along a network of mm to ~10 cm thick fracturesand shear zones [e.g., Leguéré, 1976; Lahondère et al., 2012; Quesnel et al., 2016a]. In contrast, serpentinizationis pervasive along the base of the nappe, forming a “sole” with an intense internal deformation [Avias, 1967;Orloff, 1968; Guillon, 1975; Leguéré, 1976; Cluzel et al., 2012a; Quesnel et al., 2013, 2016a]. The thickness of thissole is a few tens of meters in the Massif du Sud but reaches a few hundred meters in the northwesternklippes [Guillon, 1975; Maurizot et al., 2002]. Deformation along the serpentine sole is commonly interpretedas resulting from the emplacement of the nappe [e.g., Cluzel et al., 2012a]. Alternatively, deformation alongthe sole may partly result from subsequent extensional reactivation of the basal contact of the nappe [seeLagabrielle and Chauvet, 2008, Figure 9].

A general consensus exists that the Peridotite Nappe originated from the Loyalty Basin (Figure 1) [e.g., Avias,1967; Guillon, 1975; Collot et al., 1987; Cluzel et al., 2012a] so that it must have been emplaced with aroughly southwest vergence. However, kinematic data are scarce [Guillon and Routhier, 1971; Guillon, 1975;Poutchovsky and Récy, 1982] and mostly emanate from the rock units underneath the nappe [e.g., Tissotand Noesmoen, 1958; Gonord, 1977; Paris, 1981; Cluzel et al., 2001; Maurizot, 2011]. Quesnel et al. [2013]presented field evidence for top-to-SW shearing along the serpentine sole of the Koniambo Massif butleft it open as to whether this deformation relates to the emplacement of the nappe or to later extensionalreactivation.

The Peridotite Nappe includes a number of felsic dykes, some with boninitic affinity, dated at ~53Ma [Cluzelet al., 2006]. These dykes probably reflect a stage of high-temperature intraoceanic subduction soon after theinversion of a mid-oceanic spreading ridge [Ulrich et al., 2010; Cluzel et al., 2012a]. Right beneath thePeridotite Nappe, local slivers of metabasic rocks preserve evidence of a high-temperature amphibolite faciesevent dated at ~56Ma [Cluzel et al., 2012b]. These slivers, interpreted as relics of a metamorphic sole, supportthe hypothesis of a warm intraoceanic subduction at ~53–56Ma. As a result, the emplacement of thePeridotite Nappe onto the Norfolk Ridge microcontinent is likely younger than ~53Ma. The emplacementof the nappe occurred after that of the underlying Poya Terrane sheeted complex, as suggested by thepresence of clasts derived from the latter and the basement, but not from the peridotites, within theEocene foreland basins [Cluzel et al., 2001, 2012a]. Pelagic sedimentation persisted until the mid-Eocene inthe Poya Terrane [Cluzel et al., 2012b]. Therefore, the Peridotite Nappe was probably emplaced later than~45Ma. An even tighter time constraint is provided by the age of the uppermost sediments buried beneatheither the Poya sheeted complex or the Peridotite Nappe itself. According to biostratigraphic constraints, theoldest possible age of these uppermost sediments is ~38Ma at Nepoui and ~34–35Ma near Bourail andaround Noumea [Cluzel et al., 2001; Maurizot and Cluzel, 2014] (see Figure 1 for location). Close to Noumea,the Saint-Louis granodiorite, a small intrusion dated at ~27Ma, cuts across the basal contact of thePeridotite Nappe [Paquette and Cluzel, 2007]. Hence, near the southern end of the Grande Terre, the nappewas emplaced at some time between ~35 and ~27Ma.

Following its exposure to aerial conditions, the Peridotite Nappe was subjected to intense weathering underwarm andwet climatic conditions, which led to the development of thick laterites. Leaching of the peridotitesby meteoric waters induced a redistribution of elements, leading to nickel mineralization at the base of theweathering profile [e.g., Orloff, 1968; Leguéré, 1976; Paris, 1981]. World class nickel ore deposits have beenexploited in New Caledonia for more than a century. Most observations and measurements reported in thiswork were collected on two large mining sites, in the Koniambo Massif (as detailed in Quesnel et al. [2016a])and in the nearby Kopeto-Boulinda Massif (Figure 1). Magnesite veins, which frequently occur along theserpentine sole of the Peridotite Nappe, represent another by-product of the leaching of the peridotites bymeteoric fluids [Ulrich, 2010; Quesnel et al., 2013, 2016b]. In the KoniamboMassif, Quesnel et al. [2013] showedthat at least some of these veins were emplaced during shearing along the sole.

2.3. The Offshore Domain

The Grande Terre is surrounded by a ~5 to 15 km wide reef lagoon system. Beyond the fringing reef, theisland is flanked by two deep basins (Figure 1).

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 4

Page 12: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

To the northeast, the Loyalty Basin is separated from the Grande Terre by a steep submarine slope that hasseveral northeast dipping normal faults of Neogene age [Bitoun and Récy, 1982; Chardon and Chevillotte,2006; Chardon et al., 2008]. Based on geophysical data, the Loyalty Basin is floored by oceanic crust and repre-sents the basin from which the Peridotite Nappe originated [Bitoun and Récy, 1982; Collot et al., 1987]. Thiscrust is overlain by a thick sedimentary sequence. The lower part has geometrical relations indicating deposi-tion soon after the emplacement of the Peridotite Nappe whereas the upper part relates to Neogene exten-sional tectonics [Bitoun and Récy, 1982; Chardon and Chevillotte, 2006]. On a NW-SE trending cross sectionalong the axis of the Loyalty Basin, Bitoun and Récy [1982] showed that the lower sedimentary sequence is~3.5 to 4 km thick in the part of the basin facing the Grande Terre and only ~0.5 km thick further northwest,which suggests that these sediments essentially consist of erosional products derived from the Grande Terre.

To the southwest, the New Caledonia Basin is separated from the Grande Terre by another steep submarineslope that includes southwest dipping normal faults of Neogene age [Rigolot, 1989; Chardon and Chevillotte,2006]. Hence, the Grande Terre appears as a large horst since at least the late Miocene [e.g., Chardon andChevillotte, 2006; Lagabrielle and Chauvet, 2008]. The northern part of the New Caledonia Basin, facing theGrande Terre, has a fairly constant water depth of ~3500m. It is made up of a crust ~10 km thick overlainby a sedimentary sequence that defines an asymmetric basin deepening toward the island [Dubois et al.,1974; Lafoy et al., 2004, 2005; Klingelhoefer et al., 2007; Collot et al., 2008]. On seismic lines (e.g., the ZoNéCo11-01A line; see Figure 1 for location), the sediments are horizontal, showing a progressive onlap onto aprominent northeast dipping reflector. Based on long-distance correlation of borehole data, this reflectorcoincides with a major unconformity formed at some time between ~45 and ~30Ma. As discussed byCollot et al. [2008], the most likely interpretation of these features is that obduction on the Grande Terretriggered rapid subsidence and tilting of the foreland domain, which then became progressively filled withposttilt sediments. Since Dubois et al. [1974], a debate exists as to whether or not significant lithosphericconvergence occurred across the boundary between the Grande Terre and the New Caledonia Basin. Onehypothesis is that this boundary coincides with a subduction trench that was already active when obductionoccurred on the Grande Terre [e.g., Dubois et al., 1974; Rawling and Lister, 1999]. In contrast, recent interpreta-tions favor an activation of the boundary after the obduction event [e.g., Cluzel et al., 2001]. In these interpre-tations, the amount of convergence is either small (limited thrusting of the Grande Terre over the NewCaledonia Basin [Lafoy et al., 2004; Collot et al., 2008]) or large (subduction of the easternmost >100 km ofthe New Caledonia Basin beneath the Grande Terre [Paquette and Cluzel, 2007; Cluzel et al., 2012a]).Southwest vergent thrusts and folds deform the sedimentary pile along the foot of the steep submarine slopewest of the island [Rigolot, 1989; Lafoy et al., 2004], but it is unclear whether these structures reflect litho-spheric convergence or only local contraction at the front of large gravitational slides [Lafoy et al., 2004;Chardon and Chevillotte, 2006]. Further southwest, underneath the mid-Cenozoic unconformity, a structurewith half grabens up to ~40 km wide has been identified along some seismic lines, supporting the view thatthe northern part of the New Caledonia Basin is floored by thinned continental crust [Lafoy et al., 2004, 2005;Paquette and Cluzel, 2007].

3. Previous Models for the Emplacement of the Peridotite Nappe

Several authors have pointed out that the emplacement of the Peridotite Nappe and the development ofHP-LT metamorphism in the northern part of the Grande Terre were broadly contemporaneous, suggestinga genetic link between the two processes [e.g., Coleman, 1967; Brothers, 1974]. However, since the publicationof the white mica 39Ar-40Ar ages from the HP-LT domain at ~35–40Ma, interpreted as cooling ages [Ghentet al., 1994; Baldwin et al., 2007], and the dating of the topmost sediments beneath the Peridotite Nappe at~34–35Ma [Cluzel et al., 2001], a consensus has arisen that the nappe was emplaced while the metamorphicrocks were on their way to the surface. This may be viewed as a paradox and existing scenarios of theCenozoic evolution of New Caledonia differ in the way this apparent paradox is solved.

For Spandler et al. [2005], the metamorphic rocks were exhumed during a short-lived (from <44 to ~34Ma)episode of plate divergence and lithospheric extension that temporarily interrupted the process of plateconvergence and obduction. Considering the metamorphic domain, Cluzel et al. [1995] and Rawling andLister [1999] have argued for the existence of “normal-sense” ductile shear zones associated with extensionaltectonics. However, the same authors emphasize that these shear zones are deformed by tight upright to

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 5

Page 13: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

steep folds [see alsoMaurizot et al., 1989], so deducing their tectonic origin from their orientationmay be pro-blematic. Cluzel et al. [1995] and Clarke et al. [1997] reported that some of the shear zones are associated withthe superposition of lower grade rocks (from the high-grade part of the Diahot Terrane) onto much higher-grade rocks (from the Pouebo eclogitic mélange), providing stronger evidence for significant normal-sensedisplacements. According to Clarke et al. [1997], the related shearing event occurred at pressures greater than~13 kbar, equivalent to depths greater than ~45 km. Normal-sense shearing at such depths suggests a processof exhumation taking place along the subducting slab and does not clearly support the hypothesis of a briefepisode of whole lithosphere extension. Steep normal faults are also present in the high-grade part of themetamorphic domain [Cluzel et al., 1995; Rawling and Lister, 1999], but these could be late structures postdat-ing the emplacement of the Peridotite Nappe. For instance, they may be related to the Neogene event(s) thatshaped the Grande Terre as a horst [Chardon and Chevillotte, 2006; Lagabrielle and Chauvet, 2008].

For Baldwin et al. [2007], exhumation of the metamorphic rocks and emplacement of the Peridotite Nappewere synchronous but occurred at different sites along the strike of the plate boundary. Current remnants ofthe nappewest of themetamorphic domainwould have been carried there by amajor postobductionwrenchevent resulting in at least 150 km of dextral displacement along the strike of the island, in agreement with anearly proposal of Brothers [1974]. However, the nature of the main fault proposed to have accommodated thisdisplacement, the so-called “West Caledonian Fault,” is highly controversial. For Paris [1981] andMaurizot et al.[1985a], it is essentially a pre-Cenozoic structure that has been later reactivated as a steep thrust zone [seealso Guillon, 1975] and then as a SW dipping normal fault. For other authors, the West Caledonian Fault is justa Neogene SW dipping normal fault of limited [Cluzel et al., 2001] or greater [Lagabrielle and Chauvet, 2008]importance. In any case, a major postobduction wrench displacement along this fault seems ruled out.

A third scenario, recently put forward by Lagabrielle et al. [2013], considers that much of the emplacement ofthe Peridotite Nappe has occurred by gravitational sliding along a SW dipping slope produced by crustal-scale upward arching/doming of the underlying thrust stack. This model, also suggested by Guillon andRouthier [1971] and Cluzel et al. [1995], has the advantage that it readily solves the above paradox if the apexof the arch, from where the nappe slided away, coincides today with the higher-grade part of the meta-morphic domain. In line with this hypothesis, the highest-grade rocks occur in the same area where the mainstructural culmination of the metamorphic domain is located, close to the northern coast [e.g., Cluzel et al.,1995] (Figure 1). Gravitational sliding down a very gentle slope may have been facilitated by the presenceof the serpentine sole at the base of the nappe, acting as a weak décollement [Lagabrielle et al., 2013]. Thismodel, however, cannot be applicable to the whole of the Grande Terre. To the southeast, ignoring the effectof secondary normal faults, the basal contact of the Peridotite Nappe is horizontal around much of the Massifdu Sud and west of it until around Bourail [e.g., Gonord, 1977; Lagabrielle and Chauvet, 2008] while along thenortheastern coast the peridotites plunge abruptly toward the Loyalty Basin [Guillon, 1975; Paris, 1981].Therefore, this region seems to lack a structural culmination from which the nappe could have slippeddown. Despite this restriction, gravitational sliding could account for the geological situation in at least thenorthwestern half of the island.

Finally, although they also pointed out the synchronism between the emplacement of the Peridotite Nappeand the exhumation of the HP-LT rocks, Cluzel et al. [2001] developed a scenario that does not clearly addressthis issue [see also Cluzel et al., 2012a]. Hence, of the different tectonic models proposed so far, the oneinvolving an emplacement of the Peridotite Nappe by gravitational sliding [Lagabrielle et al., 2013] seemsthe most simple and appropriate. Alternative models implicitly assume that the nappe has been emplacedthrough horizontal contraction sustained by plate convergence [e.g., Guillon, 1975; Cluzel et al., 2001,2012a; Spandler et al., 2005]. Thus, the various models proposed to date involve different mechanisms ofemplacement for the Peridotite Nappe, and it is the aim of this paper to assess which mechanism is the mostlikely. This, in turn, may provide constraints on how obduction occurred in New Caledonia.

4. The Deformation in the Northwestern Klippes of the Peridotite Nappe4.1. The Koniambo Massif

In the Koniambo Massif, the Peridotite Nappe is exposed from its base, near sea level, up to ~800m elevation(Figures 2a and 2b). Map relations [Carroué, 1972;Maurizot et al., 2002] show that the basal contact is subhor-izontal around much of the massif. The Koniambo Massif consists of harzburgites with interlayers of dunite

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 6

Page 14: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

that define a compositional layering with a fairly regular ENE-WSW strike and a ~50° southward dip [Maurizotet al., 2002]. With regard to the degree of serpentinization, Maurizot et al. [2002] distinguished three mainrock types, namely, (i) moderately serpentinized peridotites in which the primary compositional layering ispreserved, (ii) highly serpentinized peridotites, and (iii) massive serpentinites. The latter form the serpentinesole of the nappe about 200m thick. The highly serpentinized peridotites overlie the sole and form a distinctlayer which, according to Maurizot et al. [2002], is ~200m thick on the western flank of the massif butthins out further east (Figures 2a and 2b). The moderately serpentinized peridotites occupy the higher partof the massif.

An analysis of deformations associated with distinct serpentine polymorphs and magnesite has been pre-sented by Quesnel et al. [2016a], and here we summarize the results of this work. The mineralogical contentof veins, fracture infillings, and cleavage domains has been determined through a multitechnique approachincluding Raman spectroscopy. Although rarely quoted in the literature on New Caledonia, polygonal serpen-tine was found to be widespread, a massive and matte aspect and a pale green color being its distinctivemacroscopic characteristics. We identified three structural levels, which correlate well with the rock horizonsdefined by Maurizot et al. [2002] (Figures 2a and 3).

Figure 2. (a) Geological map of the central and southern parts of the Koniambo Massif, adapted from Carroué [1972] and Maurizot et al. [2002]. (b) General crosssection of the Koniambo Massif, located in Figure 2a. (c) Report of the main fault zones along the VA1 section in the Vavouto peninsula. Figures 2b and 2c areadapted from Quesnel et al. [2016a].

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 7

Page 15: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

The peridotites of the upper structural level are crosscut by a dense network of fractures filled with one orseveral types of serpentine. Antigorite and polygonal serpentine commonly form stepped slickenfibers alongfault planes. Fault slip analysis returned two strikingly different strain ellipsoids (Figure 3). Antigorite-bearingfaults yield a highly constrictional ellipsoid with λ1, the axis of maximum stretching, lying horizontally along aWNW-ESE trend. Polygonal serpentine-bearing faults yield an ellipsoid in the flattening field with λ3, the axisof maximum shortening, lying subhorizontally along a NW-SE trend. Microstructural observations indicatethat polygonal serpentine postdates antigorite [see also Ulrich, 2010].

The serpentine sole of the Koniambo Massif is well exposed in the Vavouto peninsula (Figure 2a). Brecciasdominate while foliated mylonitic serpentinites form a distinct basal layer at least 20m thick (Figure 3).The sole has experienced pervasive noncoaxial deformation with a top-to-SW sense of shear. Polygonalserpentine and magnesite are synkinematic phases with respect to this deformation. Along the 700m longVA1 section (Figure 2c), shallow-dipping shear zones are distributed with a ~100m lengthscale. Their meanobliquity with respect to the basal contact of the nappe is ~22°.

The main part of the intermediate structural level is similar to exposures of the upper structural level, with adense network of serpentine-bearing fractures, but this rock mass is crosscut by reverse-slip shear zonesseveral meters in thickness (Figures 2b and 3). Top-to-SW shear zones are synthetic to pervasive shearingalong the serpentine sole and also involve polygonal serpentine and magnesite as synkinematic mineralphases. Therefore, although this contact could not be observed in the field, we suggest that the shear zonesroot along the roof of the serpentine sole, with the whole sole acting as a décollement (Figure 4, relationship

Figure 3. Summary of the results of structural analysis in the Koniambo Massif, adapted from Quesnel et al. [2016a]. (left) Block diagram depicting the style ofdeformation in the three structural levels identified in the massif. For the upper structural level, the drawing shows the deformation associated with polygonalserpentine. (right) Synthesis of the results of strain inversion as a function of the structural level (vertically) and as a function of the main minerals associated withdeformation, which enable a sequence of events to be identified (horizontally). Strain inversion of fault slip data has been carried out with the FaultKin software ofAllmendinger et al. [2012]. The parameter R relates to the shape of the strain ellipsoid (R = (ε2� ε3)/(ε1� ε3) hence R equates 0 for pure constriction, 0.5 for planestrain, and 1.0 for pure flattening).

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 8

Page 16: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

“r1”). In Figures 2b and 4, the roof of the sole is schematically shown as a distinct fault, but it most likelycorresponds to a zone across which the intensity of brecciation progressively diminishes. In addition, someshear zones may root at deeper level (Figure 4, relationship “r2”). The NE-SW horizontal spacing betweentwo consecutive top-to-SW shear zones, SZ1 and SZ2, is ~450m (Figures 2a and 2b). This is slightly less thanthe thickness of the upper structural level in the Koniambo Massif and much less than the thickness of thePeridotite Nappe in the Massif du Sud (≥1.5 km). This narrow spacing suggests that at least some of the shearzones do not cross the nappe up to the surface but remain confined to the intermediate structural level.Although rock exposures are abundant in the open pits on the top of the Koniambo Massif, such shear zonescould not be observed in the upper structural level. Based on this indirect evidence, we suggest that the shearzones connect to another flat-lying décollement located along the roof of the intermediate structural level(Figure 4, relationship “r3”). As for the roof of the sole, this décollement might not correspond to a distinctfault but to a zone across which the displacement along the shear zones is absorbed through diffuse faulting.

Another prominent fault zone, SZ3, is exposed in the intermediate structural level (Figures 2a, 2b and 3). Itdips southwestward and involves a ~30m thick lens of fine-grained sediments [Maurizot et al., 2002].These sediments resemble the volcano sedimentary series that immediately underlie the Peridotite Nappe[e.g., Cluzel et al., 2001]. Hence, the fault zone likely roots slightly beneath the nappe. Alternatively, it may rootalong the basal contact of the nappe since this contact commonly displays sheets of serpentinite mixed withthe volcano sedimentary rocks [e.g., Maurizot et al., 1985b]. Shear sense criteria indicate a reverse-slip displa-cement along SZ3, consistent with the incorporation of rocks from beneath the nappe into the fault zone andwith the probable offset of the roof of the serpentine sole across it (Figures 2a and 2b). This top-to-ENE faultcan be viewed as a conjugate shear with respect to the top-to-SW SZ1 and SZ2 shear zones (Figure 3).

The upper structural level provides a record of a temporal change from WNW-ESE horizontal stretching,during antigorite crystallization, to NW-SE horizontal shortening, during polygonal serpentine crystallization.The origin of the other variations in orientation of the principal strain axes is more difficult to assess becausethey involve the same serpentine polymorph (polygonal serpentine) and occur across distinct levels of thenappe (Figure 3). The change from NW-SE shortening in the upper structural level to NE-SW shearing in

Figure 4. Schematic cross section depicting the vertical distribution of deformation associated with serpentinization in the KoniamboMassif (left; from Quesnel et al.[2016a]) and the Kopéto-Boulinda Massif (right; this work). r1 to r3 are geometric relationships discussed in the text. The thickness and the topography of the nappeare poorly constrained and are likely to have changed during deformation.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 9

Page 17: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the serpentine sole could reflect a spatial rather than a temporal evolution, i.e., vertical strain partitioning at aspecific evolutionary stage of the nappe. The ~N-S direction of shortening recorded at an intermediateelevation, on roadsite “620m” (Figures 2a and 3), may support this view. Despite this, it can be argued thatthe observed changes in shortening direction are more likely to reflect a temporal evolution [Quesnel et al.,2016a]. The main argument is that it seems difficult to conceive a tectonic setting in which the nappe wouldundergo horizontal shortening at right angle to its direction of displacement (assuming that the latter isgiven by the direction of shear along the basal décollement). The observed sequence of minerals associatedwith deformation, from antigorite to polygonal serpentine�magnesite, is consistent with progressive rockcooling [Ulrich, 2010], which may reflect a change of the geothermal gradient and/or the progressive reduc-tion in thickness of the nappe.

4.2. The Kopéto-Boulinda Massif

The geology of the Kopéto-Boulinda Massif is broadly similar to that of the Koniambo Massif. The basalcontact of the Peridotite Nappe is subhorizontal as well and underlined by a serpentine sole that thickenssouthwestward [Carroué and Espirat, 1967; Carroué, 1972; Maurizot, 2007]. In the northern part of the massif,the most remarkable feature is a ~250m thick sheet of serpentinite running along the southern flank of the

Figure 5. Geological map of the northern part of the Kopéto-Boulinda Massif, adapted from Maurizot [2007]. Structural data according to this work.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 10

Page 18: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Kopéto Range (Figure 5). Guillon [1975] interpreted this structure as a major south dipping normal fault. Incontrast, Maurizot et al. [1985b] reported the sheet as dipping to the north at an angle around 45°. To thenortheast, the sheet seems to root into the serpentine sole, while to the west, map relations suggest thatthe southern margin of the sheet is a fault rooted farther north, slightly beneath the basal contact of thePeridotite Nappe [Maurizot et al., 1985b; Maurizot, 2007] (Figure 5). On the southwestern flank of theKopéto Range, the serpentine sole stands at elevations around 200m. Further east, the serpentinite sheetclimbs among the peridotites up to the main lateritic cover of the massif, at ~500m elevation.

Within the serpentinite sheet, rocks show a pronounced planar fabric underlined by a disjunctive anastomos-ing cleavage (Figure 6). In the area of higher elevations, where good exposures are provided by the accessroad to the Kopéto mining site, this foliation dips about 70° to the north (Figure 5, insert). Map contourssuggest that the serpentinite sheet as a whole is steeply dipping (Figures 5 and 6g). The sheet includes agreat number of shear planes with a mean ENE-WSW strike and variable northward dips. They bear striationswith a mean NNW-SSE trend (Figure 5, insert). On a 10 cm to ~5m scale, sigmoidal cleavage domains occur in

Figure 6. (a–f) Field views inside the South Kopéto Shear Zone (from the area shown as inset in Figure 5). (g) Sketch depicting the orientation and relations betweensecondary shear zones within the South Kopéto Shear Zone.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 11

Page 19: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

between parallel shear planes, defining local shear zones that consistently indicate a top-to-S/SE sense ofshear (Figure 6). Most shear planes have dips between ~35° and ~55° (Figures 5, 6b, 6c, and 6e), but somehave dips as low as ~10° (Figures 6d and 6f). A large shallow-dipping shear zone marks the southern marginof the serpentinite sheet along the mine access road (Figure 6a). The bulk picture (Figure 6g) is consistentwith the serpentinite sheet representing a major shear zone and the local shear zones representing asso-ciated C0-type shear bands [cf. Berthé et al., 1979; see also, e.g., Passchier and Trouw, 1996]. Hereafter, the ser-pentinite sheet will be referred to as the South Kopéto Shear Zone (SKSZ). Some of the smaller shallowdipping shear zones could have been formed lately during the development of the SKSZ (Figure 6g, case “a”).

Further south, smaller-size structures equivalent to the SKSZ are found (Figure 7). Associated striations havethe same orientation as in the SKSZ (Figure 5). Figure 7a illustrates the case of reverse-slip shear planes rootedinto a flat-lying shear zone along which ultramylonitic serpentinites with fish-shaped clasts of tremolite docu-ment a top-to-SSW sense of shear (Figure 7b). Other sites display shear zones a few meters (Figure 7e) to afew tens of meters thick (Figure 7c). These shear zones have intermediate to steep northward dips andconsistent top-to-SSE kinematics (Figures 5, 7c, 7d, and 7e).

Some of the shear zones contain one or several dismembered felsic veins (e.g., the white layer offset by anoblique shear band at the core of the shear zone in Figure 7e). The vein(s) and the host shear zone are com-monly subparallel. In at least some cases, this parallelism does not result from the transposition of the veindue to deformation; rather it reflects the fact that the shear zone was formed parallel to the vein. This indi-cates that the distribution of the shear zones may have been influenced by the presence of a network of felsicveins. The veins are intimately associated with tremolite� chlorite mineralizations [Lahondère et al., 2012]. Inzones of lesser strain, tremolite locally occurs as sheaves of large crystals with no preferred orientation; inzones of higher strain, tremolite occurs as clasts (Figure 7b). Together with the dismembered aspect of thefelsic veins, this indicates that much of the activity of the shear zones, if not all of it, postdates the emplace-ment of the veins and associated alteration. No analytical work has been carried out to determine the natureof the serpentine polymorph(s) involved in the shear zones. Nevertheless, judging from the pale green colorof cleavage domains within some of the shear zones (Figures 6b, 7b, and 7e), and by analogy with the situa-tion in the Koniambo Massif [Quesnel et al., 2016a], we hypothesize that polygonal serpentine is again thedominant synkinematic mineral phase.

5. Additional Constraints From the Central Part of Northern New Caledonia

In addition to the Massif du Sud and the northwestern klippes, many exposures of ultramafic rocks are foundalong the axis of the Grande Terre (Figure 1). The smaller exposures are described as discontinuous sheets ofserpentinite spread along fault zones [e.g., Gonord, 1977; Paris, 1981] while the larger bodies include moder-ately serpentinized peridotites. There has beenmuch debate about the significance and initial structural posi-tion of these rocks [e.g., Brothers, 1974; Gonord, 1977; Paris, 1981; Maurizot et al., 1985b], but there is broadagreement that the larger bodies represent remnants of the Peridotite Nappe. A clear example is providedby the Tchingou Massif (Figure 1) which overlies pre-Coniacian basement rocks through a subhorizontal con-tact [e.g., Guillon, 1975; Maurizot et al., 1985b].

In map view, most ultramafic rocks close to the axis of the island appear as narrow strips with a NW-SE to W-Estrike. This is the case in the area around the Ougne summit (Figure 8a), near the northwestern tip of theGrande Terre (Figure 1). There, the ultramafic rocks and closely associated mafic rocks have been interpretedeither as fragments of an ophiolite that would have initially underlain the Late Cretaceous-Eocene sedimen-tary series of the same area [Brothers, 1974; Maurizot et al., 1989] or as remnants of the Peridotite Nappe andunderlying Poya Terrane thrust sheets [Cluzel et al., 1995; Maurizot, 2011]. The map of Figure 8a is mainlybased on Maurizot et al. [1989] and satellite images. A composite cross section has been derived from thismap (Figure 8b). The ultramafic rocks, which consist of partly serpentinized peridotites overlying a serpentinesole, form the core of a series of synclines. This strongly supports the view that the ultramafic rocks do notoriginate from beneath the sedimentary cover but represent remnants of the overlying Peridotite Nappe.The fold wavelength is approximately 750m, in contrast with the nearly flat attitude, at the same scale, ofthe basal contact of the Peridotite Nappe in the northwestern klippes (Figures 1, 2a, 2b, and 5).

The folds occur in a domain where the Late Cretaceous-Eocene sediments are pervasively schistosed andunderwent metamorphism at temperatures around 270°C [Vitale Brovarone and Agard, 2013]. Schistosity

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 12

Page 20: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

planes dip steeply northeastward as part of the fan-like structure reported in Figure 1 and section 2.1. Thisschistosity is parallel to the axial plane of tight overturned folds (Figure 8c). The large synclines are also asym-metric with northeast dipping axial surfaces, and additionally, a northeast dipping thrust cuts across thestructure (Figure 8b). Thus, top-to-SW shearing is recorded in this area, in agreement with descriptions of

Figure 7. Field views of subsidiary shear zones in the footwall of the South Kopéto Shear Zone (location in Figure 5).

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 13

Page 21: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Figure 8. (a) Geological map of the Ougne summit area (location in Figure 1) based onMaurizot et al. [1989], the updated version of Maurizot et al.’s map available athttp://explorateur-carto.georep.nc (which uses the stratigraphic scheme of Cenozoic sediments revised by Maurizot [2011]), satellite images, and field observationscarried out on the southern flank of the Ougne summit. (b) Composite cross section located in Figure 8a, deduced from the same data set except the fieldobservations. (c) Field view of tight overturned folds in schistose metasediments. (d) Modified cross section of the Ougne summit taking into account the fieldobservations reported in Figure 8a. The same type of geometry likely exists all along cross section (Figure 8b) but is not drawn because field data are not available.(e) Subvertical schistosity in a serpentinite sheet within the metasediments.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 14

Page 22: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the wider region given by Cluzel et al. [1995] andMaurizot [2011]. Field observations on the southern flank ofthe Ougne summit further document that, away from the contact with the overlying peridotites, the serpen-tine sole of the Peridotite Nappe is steeply schistosed (Figure 8e) and interdigitated with the underlyingmetasediments (Figure 8a). Figure 8d shows a modified cross section of the Ougne summit according tothese observations. The geometry of the klippe is strikingly similar to that proposed by Maurizot et al.[1985b] for the Tchingou Massif. This geometry implies the existence of a contrast in strength between theperidotites and the serpentinites, the latter being approximately as weak as the underlying metasediments.This seems to contradict the conclusion we previously drew from the Koniambo Massif, that the peridotitesoverlying the serpentine sole are enough affected by serpentinization to be nearly as weak as pure serpenti-nite [Quesnel et al., 2016a]. The degree of serpentinization in the peridotites of the Ougne summit area is notknown; therefore, the contradiction is perhaps only apparent. Regardless, not only has the serpentine soleundergone~NE-SW shortening, the overlying folded and faulted peridotites have as well. Therefore, at leastthe lower levels of the Peridotite Nappe were involved in compressional deformation. Because of the largecompetence contrast between the peridotites and the underlying rocks implied by the geometry inFigure 8d, the thickness of the folded layer is likely to be less than the fold wavelength [e.g., Ramsay andHubert, 1987]. Hence, the folded layer is probably less than ~750m thick, and this may represent what wasleft of the Peridotite Nappe when folding occurred.

6. Implications for the Mechanism of Emplacement of the Nappe6.1. The Record of Compressional Deformation in the Northwestern Klippes

Figure 4 shows schematically how major shear zones are distributed in the Koniambo Massif [Quesnel et al.,2016a] and in the northern part of the Kopéto-Boulinda Massif according to the observations reported insection 4.2.

The serpentine sole represents a zone of strong tangential shear at the base of the Peridotite Nappe. As sta-ted in section 2.2, this deformation may reflect the emplacement of the nappe [e.g., Cluzel et al., 2012a]and/or a later episode of extensional reactivation [Lagabrielle and Chauvet, 2008]. The top-to-SW sense ofshear observed in the Koniambo Massif is consistent with both interpretations: in the first case because thereis a consensus that the nappe originates from the Loyalty Basin and in the second case because it may beassumed that the basal contact of the nappe was reactivated as a southwest dipping detachment. The pre-sence of localized normal-sense shear zones such as those visible along the VA1 section (Figure 2c) might beviewed as supporting the hypothesis of an extensional setting for at least part of the deformation recordedby the sole. However, if these shears represent normal fault zones related to a late extensional event, theirupward extension into the rock mass above the sole should occasionally be visible, taking into account theshort spacing (~100m) between the shear zones. This is not the case from our experience; therefore, thenormal-sense shear zones seem to be confined to the serpentine sole [see also Lahondère et al., 2012,Figure 115]. As a result, the shear zones of Figure 2c are better interpreted as subsidiary C0-type shear bandsdeveloped within a ~200m thick zone of tangential shear (Figures 3 and 4). Their mean obliquity of ~22° withrespect to the basal contact of the nappe is consistent with this interpretation [e.g., Passchier andTrouw, 1996].

According to our observations, all the shear zones found above the serpentine sole are reverse-slip shears inthe present state. It may be asked whether this represents their original attitude or whether they were oncesignificantly tilted. For instance, in the hypothesis where the Peridotite Nappe has been emplaced as a largegravitational slide (see section 3), the slide is supposed to have occurred on a southwest dipping slope;therefore, the present subhorizontal attitude of the basal contact of the nappe should be restored into asouthwestward dip in the initial state. Provided that the shear zones developed during sliding, they shouldbe back tilted as well, perhaps up to the point that their original attitude was that of normal-sense shearzones. This hypothesis is tested below.

In the case of the Koniambo Massif, the presently northeast dipping shear zones (SZ1 and SZ2; Figures 2band 3) keep a reverse sense of slip unless back tilting amounts to more than ~25°, far more than the topo-graphic slope of ~3° proposed in the gravitational slide hypothesis [Lagabrielle et al., 2013]. For the presentlysouthwest dipping shear zone (SZ3), back tilting would only strengthen its reverse slip nature. The sedimentspinched in this shear zone also attest for its origin as a thrust. As a consequence, the fact that SZ3 crosscuts

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 15

Page 23: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

and offsets the serpentine sole further weakens the hypothesis that part of the internal deformation of thesole may result from a younger extensional event.

In the case of the Kopéto-Boulinda Massif, the SKSZ has a northward dip of about 70° at a structural height~300m above the serpentine sole (Figures 5 and 6g). It is possible that the shear zone steepens even moreupward, in a section of the nappe now eroded (Figure 4). Back tilting the whole nappe would make the SKSZless steep but would maintain its reverse-slip kinematics. This is also the case for the subsidiary steep shearzones located further south (Figures 7c–7e). In addition, the inclusion within lower levels of the nappe of asheet of volcano sedimentary rocks belonging to the substratum, as suggested by map contours west ofthe Kopéto Range (Figure 5), supports the view that the southern margin of the SKSZ is a thrust fault in origin.This fault roots northward [see also Maurizot et al., 1985b] while top-to-south shearing is recorded across theSKSZ (Figure 6). Therefore, the two structures probably relate to the same tectonic event.

In summary, the reverse-slip shear zones of the Koniambo and Kopéto-Boulinda klippes are true compres-sional structures in origin. In the Koniambo Massif, these shear zones accommodate NE-SW shortening,consistent with the top-to-SW sense of shear recorded by the serpentine sole (Figures 3 and 4). In theKopéto-Boulinda Massif, the SKSZ strikes ~W-E and accommodates top-to-SSE shearing, which implies asmall component of dextral displacement. The steepest cleavage domains within the SKSZ further includea set of west plunging striations implying a more oblique-slip displacement. This, together with the verysteep attitude of the shear zone, suggests that it has been formed as an oblique thrust with a dextral wrenchcomponent (Figure 4). Some authors already pointed out the influence of dextral transcurrent tectonics in theearly Cenozoic evolution of the Grande Terre [Gonord et al., 1973; Gonord, 1977; Paris, 1981; Cluzel et al., 2001;Titus et al., 2011].

Top-to-SSE shearing along the SKSZ is fairly consistent with the deformation recorded by polygonalserpentine-bearing faults in the upper structural level of the Koniambo Massif (Figure 3). The “strike-slip”regime seen in Figure 3, with both λ1 and λ3 having low plunges, is dominated by NNE-SSW strikingreverse-sinistral faults [Quesnel et al., 2016a]. These faults may represent conjugate shears with respect tothe SKSZ. The flattening-type strain ellipsoid returned by the fault slip analysis implies stretching along thesteep λ2 axis, which is consistent with the hypothesis of a transpressional setting [e.g., Fossen, 2010]. Thesefeatures suggest that deformation progressively evolved from a transpressional setting along a ~W-E struc-tural trend (recorded along the SKSZ and in the upper structural level of the Koniambo Massif) to pure NE-SWcompression (recorded in the serpentine sole and the intermediate structural level of the Koniambo Massifand possibly as local shear zones in the Kopéto-Boulinda Massif; cf. Figures 5, 7a, and 7b).

The observation of compressional structures in the Koniambo and Kopéto-Boulinda klippes contrasts withthe opinion of Lagabrielle et al. [2013] that no such structure exists in the Peridotite Nappe. Lagabrielleet al. [2013] used the apparent lack of compressional structures as a key argument for an emplacement ofthe nappe through gravitational sliding. Conversely, the presence of compressional structures is consistentwith the alternative hypothesis that the nappe has been emplaced through horizontal contraction sustainedby plate convergence. Nevertheless, natural examples of gravitational slides (e.g., along a salt or shaledécollement down the slope of a passive continental margin) and analogue models of the process also showthe development of compressional structures in the frontal part of the nappe [e.g., Demercian et al., 1993;Morley and Guerin, 1996; Fort et al., 2004; Loncke et al., 2006; Mourgues et al., 2009; de Vera et al., 2010; Brunand Fort, 2011]. In the next section, we examine whether compressional deformation in the Koniambo andKopéto-Boulinda klippes could reflect the latter case.

6.2. Compression at the Front of a Large Gravitational Slide?

Natural and analogue examples of gravitational slides display compressional structures around the toe ofthe topographic slope. Compression tends to propagate upslope during continued sliding [Fort et al., 2004;Brun and Fort, 2011] but always remains confined to the lower half of the slope [e.g., Morley and Guerin,1996; de Vera et al., 2010]. Therefore, a key question is whether the rocks of the Koniambo and Kopéto-Boulinda klippes were located near the toe of a regional southwest dipping slope by the time the nappewas emplaced.

As stated in section 2.3, the Grande Terre is flanked by the New Caledonia Basin, which has a water depth of~3500m and a sedimentary sequence resting horizontally against a northeast dipping unconformity of mid-

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 16

Page 24: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Cenozoic age. This geometry, well visible along the ZoNéCo11-01A seismic line (Figures 1 and 9a), is inter-preted as resulting from a two-stage evolution [Collot et al., 2008]: (i) the obduction event on the GrandeTerre triggered almost instantaneous tilting of the foreland domain and (ii) the trough became progressivelyfilled with posttilt sediments. Hence, at the time Peridotite Nappe was emplaced, a trough existed immedi-ately to the southwest. Its depth can be estimated as follows. Along the ZoNéCo11-01A seismic line,the eastern end of the posttilt sedimentary wedge is ~3600m thick (Figure 9a). Further east, closer to theGrande Terre, the pile is thicker, but this may partly result from folding and thrusting, as visible along theseismic line [Lafoy et al., 2004]. On one hand, subsidence due to the sediment load may have increasedthe deepening of the eastern part of the basin. On the other hand, compaction of the sedimentary pile cer-tainly reduced its thickness. In addition, the top of the pile is at a depth of 3500m at present. The currentwater columnmay partly relate to subsidence due to Neogene extension, but this tectonic event of moderateintensity probably had a limited impact. Judging from the horizontal attitude of the posttilt deposits, posttilt-ing deepening of the basin, if any, seems to have been spatially homogeneous. As a result, by the time thePeridotite Nappe was emplaced, the trough southwest of it was most probably at least 3600m deep(Figure 9b). There is no tight constraint on elevations at that time on the Grande Terre. In some young oro-gens occurring as elongated islands, present elevations can reach values as high as ~3600m, as in NewZealand or Taiwan. In the case of the Grande Terre, the orogen is narrower and was capped by a thick sheetof dense ultramafic rocks; therefore, 3600m is a reasonable upper bound for the maximum elevations everreached in this orogen. As a consequence, the rocks of the northwestern klippes were initially located inthe upper part of the regional slope (Figure 9b). As discussed above, compressional structures are notexpected at this level of the slope in the hypothesis of a gravitational slide. Hence, the structures seen inthe Koniambo and Kopéto-Boulinda klippes seem to rule out gravitational sliding as the mechanism ofemplacement of the Peridotite Nappe.

The reasoning above assumes that the southwest dipping slope joining the Grande Terre orogenic domainto the New Caledonia Basin was fairly regular (Figure 9b). However, at present, a flat domain exists at anintermediate level of the slope, corresponding to the reef lagoon system that surrounds the island. Thismorphology partly results from the building of the fringing reef since the late Neogene. The reef musthave developed at very shallow water depths and without a strongly eroding aerial slope nearby; hence,the flat domain existed prior to the late Neogene. The lagoon is presently 5 to 10 km wide in the area of

Figure 9. (a) Cross section showing the relations between the Grande Terre and the New Caledonia Basin (location in Figure 1). (b) The same cross section at~40–35Ma, build for testing the hypothesis that the Peridotite Nappe was emplaced as a large gravitational slide.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 17

Page 25: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the Koniambo and Kopéto-Boulinda klippes (Figure 1), but the flat domain may have extended furthernortheast as suggested by the low elevation and subhorizontal attitude of the basal contact of the nappearound the klippes. Despite the regional slope toward the New Caledonia Basin, this flat domain couldhave played the role of an intermediate slope toe against which a gravitational slide might have experi-enced compressional deformation.

In Figure 9a, it is suggested that the flat domain actually results from block tilting between two southwestdipping normal fault zones associated with Neogene extension. One fault zone is the set of normal faultsreported from the steep submarine slope immediately southwest of the reef [Rigolot, 1998; Chardon andChevillotte, 2006]; the other coincides with the West Caledonian Fault (see section 3). According to Gonord[1977], the West Caledonian Fault bounds the Koniambo Massif as a network of steep NW-SE-trending faults,which is consistent with the rectilinear aspect of lithological contours and the present landscapemorphologyin this area. In Figure 2a, the probable existence of this fault zone is shown with a box ornament added to thebasal contact of the nappe. The contact stands at higher elevations there than it does elsewhere around themassif, which may result from drag folding due to normal-sense displacement along the fault zone. Hence,Neogene block tilting could perhaps account for the secondary development of a flat segment in an initiallycontinuous regional slope (Figure 9).

However, this interpretation is merely a possibility, and we cannot exclude that the flat domain alreadyexisted when the Peridotite Nappe was emplaced. For instance, if the hypothesis of a large thrust carryingthe Grande Terre over the New Caledonia Basin at the end of the obduction event is correct (seesection 2.3), then the region of the northwestern klippes likely coincided with a flat domain above this thrust.As mentioned above, this domain could have played the role of a local slope toe against which the PeridotiteNappe might have undergone compression in the context of a large gravitational slide. Therefore, in order toreach a firm conclusion, the structural record in klippes of the Peridotite Nappe located away from the flatdomain must be considered.

6.3. Spatial Variations in Structural Style Associated With Compression, Implications

In the area around the Ougne summit, the Peridotite Nappe has undergone compressional deformation(Figure 8). This area is located about 10 km at the rear of the northwestern klippes, significantly closer tothe main structural culmination of the HP-LT metamorphic domain (Figure 1). In the gravitational slidehypothesis, sliding is supposed to start from this culmination. The Ougne summit area lies fairly close to thisaxis, and as discussed in section 6.2, compression should not be observed in such a shallow part of the south-west dipping regional slope (Figure 9b). In contrast, compression within any part of the nappe is consistentwith its emplacement through crustal-scale horizontal contraction. The northwestern klippes are flat lying,overlying at high angle the steep fault complex of the Poya Terrane [Cluzel et al., 2001]. In the Ougne summitarea, the Peridotite Nappe is folded in association with the development of a steep pervasive schistosity inlow-grade metasediments. This difference in structural style implies that the Peridotite Nappe experiencedcompression at greater depth conditions toward the northeast. This feature, unexpected if the nappe hadbeen emplaced by gravitational sliding, is consistent with an emplacement through a “push from the rear”mechanism [e.g., Merle, 1986]. A push from the rear mechanism implies a context of bulk horizontal contrac-tion and plate convergence. It also implies the existence of a major thrust at the rear of the nappe, the preciselocation of which is discussed in the next section.

7. Obduction Tectonics in New Caledonia7.1. Pre-Neogene Crustal-Scale Restoration

Based on our results and on data from the literature, Figure 10a shows a crustal-scale cross section of north-ernmost New Caledonia (see Figure 1 for location) tentatively restored in a pre-Neogene setting, i.e., beforethe evolution of the Grande Terre as a large horst [e.g., Chardon and Chevillotte, 2006]. The key elements ofthis cross section are as follows.

The Peridotite Nappe originates from the Loyalty Basin [e.g., Collot et al., 1987], however, in the tectonic stageshown in the restoration (at ~34Ma), the spatial continuity of the nappe is already lost due to the surge of themetamorphic domain. To the southwest, the nappe is partly eroded and represented by a large klippe. At itsfront, the nappe is about 1 km thick, flat lying, and includes SW-vergent thrusts deduced (and projected) from

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 18

Page 26: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the Koniambo/Kopéto-Boulinda area (Figure 4). Further northeast, the nappe is about 500m thick, folded,and faulted, as deduced from the Ougne summit area (Figure 8).

The Ougne summit area is also part of the domain where schistosities in the underlying metasedimentsdefine a ~10 km wide upright fan (Figure 1). We interpret this fan as the upward extension, within the softsedimentary cover of the Norfolk Ridge, of a major thrust carrying the HP-LT rocks onto the continental base-ment of the ridge. The fan shape reflects the progressive transition from northeast dipping schistositiesresulting from thrusting to southwest dipping schistosities concordant with the southwestern limb of a largefold nappe of exhumed HP-LT rocks. A ~30 cm scale analogue of this geometry is visible in the picture of

Figure 10. (a) Crustal-scale cross section of northernmost New Caledonia (location in Figure 1) restored in a pre-Neogene setting. For the sake of simplicity, thesection omits the Poya Terrane, a thin package of highly faulted oceanic rocks between the Peridotite Nappe and the sedimentary cover of the Norfolk Ridge(for details regarding the Poya Terrane, see Cluzel et al. [2001]). See the text for a description of the section. (b) Field view illustrating the deformation of metase-diments in the footwall of the Lizard ophiolite, southwestern England (Porthleven section; see, e.g., Leveridge and Shail [2011]). This ~30 cmwide outcrop represents aplausible analogue of the crustal-scale structure of New Caledonia after obduction.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 19

Page 27: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Figure 10b, taken in metasediments from the footwall of the Lizard ophiolite, southwestern England [e.g.,Leveridge and Shail, 2011]. Figure 10a thus provides a simple explanation for the presence of the upwardschistosity fan. It differs from the interpretation of Rawling and Lister [1999] who assumed that the southwestdipping schistosities are associated with a relatively young normal-sense shear zone crosscutting the thrust-related northeast dipping schistosities. The thrust zone flanking the fold nappe likely accounts for part of theemplacement of the Peridotite Nappe having occurred through a push from the rear mechanism.

Further northeast, we propose that the fold nappe was flanked by a major northeast dipping normal-sensefault zone. Normal faulting along the northeastern coast of the Grande Terre is documented for Neogenetimes [Bitoun and Récy, 1982; Chardon and Chevillotte, 2006; Chardon et al., 2008], but these minor faultscan barely account for the huge vertical offset implied by the occurrence of the HP-LT rocks in close contactwith the Loyalty Basin peridotites which underlie the eastern lagoon according to gravity data [e.g., Collotet al., 1987; Cluzel et al., 2012a]. We suppose that, on its eastern side, the fold nappe of HP-LT rocks wasexhumed in the footwall of a northeast dipping normal fault zone and that this normal fault developed syn-chronously with the thrust zone associated with the schistosity fan (Figure 10a). This interpretation is reminis-cent of the model proposed by Chemenda et al. [1996] for Oman, which also predicts the existence of anupright fan of schistosities at the front of a fold nappe of HP-LT rocks [see also Cluzel et al., 2012a, Figure 9].

To the southwest, we accept the common view that the Norfolk Ridge is thrust over the thinned continentalcrust of the New Caledonia Basin [e.g., Cluzel et al., 2001; Lafoy et al., 2004; Collot et al., 2008]. The relatedthrusts are shown as dashed lines in Figure 10a to emphasize that there is no direct evidence for their exis-tence to date (see section 2.3). We leave it open whether convergence along the western margin of theGrande Terre ever involved the subduction of an oceanic lithosphere [Paquette and Cluzel, 2007; Cluzelet al., 2012a]. If so, this ocean would already be subducted at the stage shown in Figure 10a, but we donot mean that this was necessarily the case as early as ~34Ma. Due to its position right above the presumedmajor thrust system, the frontal part of the Peridotite Nappe may have undergone renewed tangential shearon this occasion. In line with this hypothesis, top-to-SW shearing along the serpentine sole of the KoniamboMassif involves synkinematic magnesite veins [Quesnel et al., 2013, 2016a] formed at temperatures of only~30°C [Quesnel et al., 2016b], which suggests that the thickness of the Peridotite Nappe had already beenmuch reduced when this deformation occurred.

7.2. The Record of Oblique Convergence and its Geodynamic Setting

Several features indicate that obduction in New Caledonia occurred through oblique convergence with adextral wrench component.

First, as discussed in section 6.1, deformation associated with serpentinization in the Koniambo and Kopéto-Boulinda klippes is consistent with a progressive evolution from a transpressional setting involving dextralshearing along a ~W-E structural trend to pure NE-SW compression. Further northwest, starting from thewestern margin of the Tiebaghi Massif, the klippes of the Peridotite Nappe include a prominent zone of dex-tral shear involving southwest dipping mylonitic lherzolites [Nicolas, 1989; Leblanc, 1995]. For Nicolas [1989],this shear zone represents a transform fault inherited from the preobduction oceanic accretion period.However, as discussed by Titus et al. [2011], the trend of the shear zone is difficult to reconcile with thishypothesis. Titus et al. [2011] suggested that the shear zone represents an old structure (the paleospreadingridge?) reactivated as a wrench fault after obduction. However, as it involves high-temperature mylonites, theshear zone is unlikely to postdate obduction. Instead, it may date from the early stages of obduction. In thePoum Massif, the southwest dipping foliation of the mylonitic lherzolites bears a mineral lineation with amean~N-S trend [Nicolas, 1989; Leblanc, 1995]. Hence, dextral shearing is associated with a component ofreverse shear, consistent with a transpressional setting. In Figure 10a, we tentatively show this shear zonerooting into the basal contact of the Peridotite Nappe by analogy with the situation in the Kopéto-Boulinda Massif (Figures 4 and 5).

Second, dextral shearing along a trend parallel to the Grande Terre is also recorded in the rock units under-lying the Peridotite Nappe. Several kilometer long dextral faults are reported from the region west of theMassif du Sud [Gonord et al., 1973; Gonord, 1977]. Further northwest, in the vicinity of the lawsonite isograd,lithological contours suggest the presence of large asymmetric folds consistent with dextral wrenching [Paris,1981]. East of the isograd, the Ouégoa-Bondé fold supports this interpretation (Figure 1). At the front of the

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 20

Page 28: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

upward schistosity fan, schistosityplanes dip northeastward at anglesaround 50–55° [Maurizot et al., 1989;Cluzel et al., 1995] (Figure 8c). This sug-gests that the thrust presumed tounderlie the fan is relatively steep and,thus, includes a strike-slip component(Figure 10a).

Third, oblique convergence is consistentwith several additional features of thegeology of the Grande Terre. As pointedout by Paris [1981] and Cluzel et al.[2001], the emplacement of thePeridotite Nappe was probably diachro-nous along the strike of the island,occurring earlier in the north. The bestevidence for diachronism is the fact thatthe onset of clastic sedimentation in theforeland basins occurred progressivelylater southeastward [Cluzel et al.,2012a; Maurizot and Cluzel, 2014]. Inaddition, only the northwestern halfof the island includes a domain ofCenozoic HP-LT rocks (Figure 1). Thesetwo features are best explained byassuming that the axis of the NorfolkRidge and the zone of intraoceanic sub-duction that eventually led to theobduction of the Peridotite Nappe werenot parallel, as in most geodynamicmodels of the region [e.g., Crawfordet al., 2003; Titus et al., 2011], but signifi-

cantly oblique [Cluzel et al., 2001]. Figure 11a shows a possible configuration for this obliquity. The W-E strikeof the trench is deduced from (i) the fact that subduction probably initiated through the inversion of a mid-oceanic ridge [Crawford et al., 2003; Ulrich et al., 2010; Cluzel et al., 2012a, 2012b] and (ii) the various structuralelements of the Peridotite Nappe documenting a ~N-S direction of spreading at this ridge [Prinzhofer et al.,1980], although Titus et al. [2011] have questioned the validity of this inference. The Peridotite Nappe isnow considered as essentially a fore-arc ophiolite that recorded widespread partial melting during earlyintraoceanic subduction [Marchesi et al., 2009; Ulrich et al., 2010; Pirard et al., 2013]. As a result, it may be askedwhether the structural elements measured by Prinzhofer et al. [1980] relate to the initial mid-oceanic ridge orto later fore-arc spreading. However, the strike of the trench would also be ~W-E in the latter case becausefore-arc spreading is expected to occur at high angle to the trench [e.g., Stern et al., 2012]. The geodynamicsetting in Figure 11a implies that the northern tip of the Norfolk Ridge met the trench first. The progressiveentrance of more southerly parts of the continental ridge at the trench is consistent with the sedimentaryrecord in foreland basins and may have had two effects. It probably made the subduction of this low-densitymaterial more difficult so that only the northern margin of the ridge could be buried to HP-LT conditions. Inaddition, it may have forced the trench to drape against the ridge so that the convergent front rotated clock-wise and started to accommodate a dextral wrench component (Figure 11a), in line with the observationsreported above. As shown in Figure 11b, not only the trench but also the rocks in its hangingwall areexpected to have rotated clockwise. Therefore, the question arises whether it makes sense to use the orienta-tion of preobduction structures [Prinzhofer et al., 1980] to infer the initial strike of the trench. Prinzhofer et al.[1980] deduced a ~N-S direction of spreading from data gathered in theMassif du Sud. In contrast, their mea-surements of high-temperature mineral lineations in the northwestern klippes have amean NE-SW trend [see

Figure 11. (a) Description, in map view, of the geodynamic setting thatled to obduction of the Peridotite Nappe under dextral oblique conver-gence. (b and c) Sketches aiming at discussing the relations betweenclockwise rotation of the Peridotite Nappe, as expected from the scenarioin Figure 11a, and the orientation of preobduction structures. See the textfor explanation.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 21

Page 29: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

also Nicolas, 1989], consistent with the rotation predicted in Figure 11b. In Figure 11c, we propose that therocks of the Massif du Sud experienced no rotation because this part of the Peridotite Nappe has beenemplaced along a~N-S dextral tear fault. Following Titus et al. [2011], such a fault may exist slightly off thewestern margin of the Massif du Sud, representing a reactivation of the preobduction transform fault zoneidentified in the Bogota peninsula [Nicolas, 1989] (see Figure 1 for location). Further east, within the Massifdu Sud, Guillon and Routhier [1971] identified an important ~N-S dextral fault associated with serpentinites,the Kouakoué fault zone, which could be subsidiary to the main tear fault.

Cluzel et al. [2001, 2012a] have already proposed a sketch resembling that of Figure 11a. The main differencesare the strike of the trench before obduction, supposed to be ~NW-SE, and the initial trend of the NorfolkRidge at the latitude of the Grande Terre, taken as ~N-S, which coincides with the present orientation ofthe ridge south of the Grande Terre. To the north, the ridge would have rotated anticlockwise as a result ofits arrival at the trench, becoming parallel to the trench. Hence, the current bend, in map view, of theNorfolk Ridge would be a consequence of the obduction process [Cluzel et al., 2001] whereas we assume inFigure 11a that this bend existed before. The difference between the two models is only minor as both pre-dict a southward propagation of deformation along the Grande Terre and oblique convergence with a dextralwrench component during obduction (in Cluzel et al.’s model, dextral shearing is a necessary consequence ofthe anticlockwise rotation of the ridge). The advantage of our model is that it better integrates the data ofPrinzhofer et al. [1980]. Moreover, according to some paleomagnetic data [Ali and Aitchison, 2002], the base-ment of the Grande Terre underwent no significant rotation during the obduction. This is consistent with ourmodel in which only the trench and its hangingwall (i.e., the Peridotite Nappe) rotate (Figures 11a and 11b).

7.3. Synconvergence Exhumation of the HP-LT Rocks

The scenario of Figure 11a, characterized by a diachronous emplacement of the Peridotite Nappe along thestrike of the Grande Terre, may provide a simple explanation for the apparent paradox of the nappe havingbeen emplaced at the same time the HP-LT rocks were on their way to the surface (see section 3). In the north,the nappe could have been emplaced at ~42–40Ma, accompanied by the burial of the northern margin ofthe Norfolk Ridge. At ~36Ma, the nappe had not yet reached the south of the Grande Terre (Figure 11a),but its northern part, emplaced earlier, could have already been affected by the uplift of the HP-LTfold nappe, which was almost completed at ~34Ma (Figure 10a). This scenario is compatible with all availabletime constraints (see section 2) except the white mica 39Ar-40Ar ages from the HP-LT domain at ~35–40Ma,provided that they represent cooling ages reflecting late stages of exhumation [Ghent et al., 1994; Baldwinet al., 2007]. The latter view is questionable because most white mica 39Ar-40Ar ages were obtained on rockswith well preserved HP-LT assemblages, such as blueschist facies metabasites. Therefore, these ages mayessentially reflect the conditions of blueschist facies metamorphism, at pressures higher than 9 kbar [e.g.,Fitzherbert et al., 2005], corresponding to depths greater than 30 km. Only the youngest 39Ar-40Ar ages, at~35–37Ma, might record the exhumation to shallower depths.

According to our interpretation, the exhumation of the HP-LT rocks occurred during ongoing convergenceand crustal-scale contraction (Figure 10a). In addition to exhumation processes that may have taken placealong the subduction “channel” [e.g., Agard et al., 2009; Agard and Vitale Brovarone, 2013; Guillot et al.,2015], two factors could have assisted unroofing. One factor is the obliquity of convergence (seesection 7.2) since oblique convergence is suspected to be more favorable to exhumation than pure conver-gence [e.g., Mann and Gordon, 1996; Boutelier and Chemenda, 2008]. The second factor is the possibility thathigh erosion rates prevailed in the area above the exhuming fold nappe. Figure 10a suggests that erosion wasmuch more efficient in this area than on the southwestern flank of the island where klippes of the PeridotiteNappe are preserved. This asymmetry in the distribution of erosion fits with the current distribution of rainfall,in the range of 1600 to 3700mm/yr along the northeastern coast versus 800–1300mm/yr along the south-western coast [Maitrepierre, 2012]. The strongly asymmetrical distribution of rainfall across the GrandeTerre is a consequence of the dominant winds coming from the east-southeast, a feature that may havealready existed ~35Ma ago. Focused erosion on the northeastern flank of the island may have helped to pro-duce a fold nappe rising through the rear part of the orogen [e.g., Chemenda et al., 1996; Willett, 1999]. Thebasins on both sides of the Grande Terre have a thick sedimentary infill broadly contemporaneous withobduction (see section 2.3); hence, a large part of this infill may consist of rocks eroded from above the foldnappe. This is especially likely in the case of the Loyalty Basin where pre-Neogene postobduction deposits are

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 22

Page 30: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

~3.5 to 4 km thick in the part of the basin facing the Grande Terre; this thickness decreasing to ~0.5 km furthernorthwest [Bitoun and Récy, 1982]. Summing up, both oblique convergence and focused erosional denuda-tion may have helped to exhume the HP-LT rocks in the form of a steep fold nappe (Figure 10a).

8. Conclusions

This study provides a better understanding of the relationships between the Peridotite Nappe and its substra-tum along a cross section through the northern part of the Grande Terre. As illustrated by the Koniambo andKopéto-Boulinda massifs, the northwestern klippes are essentially flatlying but involve reverse-slip shearzones which are true compressional structures in origin. Further northeast, closer to the structural culmina-tion of the HP-LT metamorphic domain, the Peridotite Nappe is folded in association with the developmentof a steep schistosity in low-grade metasediments. This difference in structural style indicates that thePeridotite Nappe experienced compression at greater depth conditions toward the northeast, i.e., towardits root zone, suggesting a push from the rear mechanism of emplacement. Irrespective, these features areinconsistent with an emplacement of the nappe as a gravitational slide. Instead, they imply that thePeridotite Nappe has been emplaced through horizontal contraction sustained by plate convergence.Combining this evidence with additional features such as the existence of a large upward fan of schistositiesin the area where the nappe is the most deformed, a crustal-scale cross section has been built for the periodpreceding Neogene normal faulting. It involves a fold nappe of HP-LT rocks bounded from below by a majorthrust and from above by a large normal-sense shear zone, much alike in the interpretation of Chemenda et al.[1996] for Oman.

Additional considerations suggest that obduction in New Caledonia occurred through oblique convergencewith a dextral component. Oblique convergence is expected as a result of the initial obliquity betweenthe~W-E striking subduction trench and the NW-SE trending continental ribbon of the Norfolk Ridge.Cluzel et al. [2001, 2012a], who suggested a slightly different configuration, pointed out that a significant obli-quity between the trench and the continental ribbon is the easiest way to account for the diachronous recordof compression along the strike of the Grande Terre. We have shown that this diachronismmay also solve theparadox of the Peridotite Nappe being emplaced seemingly at the same time the HP-LT rocks were on theirway to the surface. Oblique convergence together with focused erosional denudation on the northeasternflank of the Grande Terre probably helped to exhume the HP-LT rocks in the form of a steep fold nappe risingthrough the rear part of the orogen.

ReferencesAgard, P., and A. Vitale Brovarone (2013), Thermal regime of continental subduction: The record from exhumed HP-LT terranes

(New Caledonia, Oman, Corsica), Tectonophysics, 601, 206–215.Agard, P., P. Yamato, L. Jolivet, and E. Burov (2009), Exhumation of oceanic blueschists and eclogites in subduction zones: Timing and

mechanisms, Earth Sci. Rev., 92, 53–79.Ali, J. R., and J. C. Aitchison (2002), Paleomagnetic-tectonic study of the New Caledonia Koh Ophiolite and the mid-Eocene obduction of the

Poya Terrane, New Zealand J. Geol. Geophys., 45, 313–322.Allmendinger, R. W., N. Cardozo, and D. M. Fisher (2012), Structural Geology Algorithms, 289 pp., Cambridge Univ. Press, U. K.Avias, J. (1967), Overthrust structure of the main ultrabasic New Caledonian massives, Tectonophysics, 4, 531–541.Baldwin, S. L., T. Rawling, and P. G. Fitzgerald (2007), Thermochronology of the New Caledonian high-pressure terrane: Implications for

middle Tertiary plate boundary processes in the southwest Pacific, Geol. Soc. Am. Spec. Pap., 419, 117–134.Berthé, D., P. Choukroune, and D. Gapais (1979), Orientations préférentielles du quartz et orthogneissification progressive en régime

cisaillant: l’exemple du cisaillement sud-armoricain, Bull. Mineral., 102, 265–272.Bitoun, G., and J. Récy (1982), Origine et évolution du Bassin des Loyauté et de ses bordures après la mise en place de la série ophiolitique de

Nouvelle-Calédonie, in Contribution à l’étude géodynamique du Sud-Ouest Pacifique, Travaux et Documents de l’ORSTOM (Paris), vol. 147,pp. 505–539, ORSTOM, Paris.

Boutelier, D. A., and A. I. Chemenda (2008), Exhumation of UHP/LT rocks due to the local reduction of the interplate pressure: Thermo-mechanical physical modelling, Earth Planet. Sci. Lett., 271, 226–232.

Brothers, R. N. (1974), High-pressure schists in northern New Caledonia, Contrib. Mineral. Petrol., 46, 109–127.Brun, J. P., and X. Fort (2011), Salt tectonics at passive margins: Geology versus models, Mar. Petrol. Geol., 28, 1123–1145.Carroué, J. P. (1972), Carte géologique de la Nouvelle-Calédonie à l’échelle 1/50000, feuille Pouembout, map sheet and explanatory notes,

Bur. Rech. Géol. Min., France.Carroué, J. P., and J. J. Espirat (1967), Carte géologique de la Nouvelle-Calédonie à l’échelle 1/50000, feuille Poya - Plaine des Gaïacs, map

sheet and explanatory notes, Bur. Rech. Géol. Min., France.Chardon, D., and V. Chevillotte (2006), Morphotectonic evolution of the New Caledonia ridge (Pacific Southwest) from post-obduction

tectonosedimentary record, Tectonophysics, 420, 473–491.Chardon, D., J. A. Austin Jr., G. Cabioch, B. Pelletier, S. Saustrup, and F. Sage (2008), Neogene history of the northeastern New Caledonia

continental margin from multichannel reflection seismic profiles, C. R. Geosci., 340, 68–73.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 23

AcknowledgmentsThis work was financially and technicallysupported by Koniambo Nickel SAS. Wethank Pierre Rossler (Société Le Nickel)for giving access to the Kopéto miningsite. Dominique Cluzel is warmlythanked for stimulating discussions andfor having made possible our visit to theKopéto-Boulinda Massif and the Ougnesummit area. Samuel Angiboust and ananonymous reviewer are thanked fortheir constructive reviews. Detailsregarding field data are available fromthe authors upon request ([email protected]).

Page 31: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Chemenda, A. I., M. Mattauer, and A. N. Bokun (1996), Continental subduction and a mechanism for exhumation of high-pressuremetamorphic rocks: New modelling and field data from Oman, Earth Planet. Sci. Lett., 143, 173–182.

Clarke, G. L., J. C. Aitchison, and D. Cluzel (1997), Eclogites and blueschists of the Pam Peninsula, NE New Caledonia: A reappraisal, J. Petrol.,38, 843–876.

Cluzel, D., J. Aitchison, G. Clarke, S. Meffre, and C. Picard (1995), Dénudation tectonique du complexe à noyau métamorphique de hautepression d’âge tertiaire (Nord de la Nouvelle-Calédonie, Pacifique, France), données cinématiques, C. R. Acad. Sci. Paris, 321, 57–64.

Cluzel, D., J. C. Aitchison, and C. Picard (2001), Tectonic accretion and underplating of mafic terranes in the late Eocene intraoceanic fore-arcof New Caledonia (southwest Pacific): Geodynamic implications, Tectonophysics, 340, 23–59.

Cluzel, D., S. Meffre, P. Maurizot, and A. J. Crawford (2006), Earliest Eocene (53 Ma) convergence in the southwest Pacific: Evidence frompre-obduction dikes in the ophiolite of New Caledonia, Terra Nova, 18, 395–402.

Cluzel, D., P. Maurizot, J. Collot, and B. Sevin (2012a), An outline of the geology of New Caledonia; from Permian-Mesozoic southeastGondwanaland active margin to Cenozoic obduction and supergene evolution, Episodes, 35, 72–86.

Cluzel, D., F. Jourdan, S. Meffre, P. Maurizot, and S. Lesimple (2012b), The metamorphic sole of New Caledonia ophiolite:40Ar/

39Ar, U-Pb, and

geochemical evidence for subduction inception at a spreading ridge, Tectonics, 31, TC3016, doi:10.1029/2011TC003085.Coleman, R. G. (1967), Glaucophane schists from California and New Caledonia, Tectonophysics, 4, 479–498.Coleman, R. G. (1971), Plate tectonic emplacement of upper mantle peridotites along continental edges, J. Geophys. Res., 76, 1212–1222,

doi:10.1029/JB076i005p01212.Collot, J. Y., A. Malahoff, J. Récy, G. Latham, and F. Missègue (1987), Overthrust emplacement of New Caledonia ophiolite: Geophysical

evidence, Tectonics, 6, 215–232, doi:10.1029/TC006i003p00215.Collot, J., L. Géli, Y. Lafoy, R. Vially, D. Cluzel, F. Klingelhoefer, and H. Nouzé (2008), Tectonic history of northern New Caledonia Basin from

deep offshore seismic reflection: Relation to late Eocene obduction in New Caledonia, southwest Pacific, Tectonics, 27, TC6006,doi:10.1029/2008TC002263.

Crawford, A. J., S. Meffre, and P. A. Symonds (2003), 120 to 0 Ma tectonic evolution of the southwest Pacific and analogous geologicalevolution of the 600 to 220 Ma Tasman Fold Belt System, Geol. Soc. Australia Spec. Publ., 22, 377–397.

de Vera, J., P. Granado, and K. McClay (2010), Structural evolution of the Orange Basin gravity-driven system, offshore Namibia, Mar. Petrol.Geol., 27, 223–237.

Demercian, S., P. Szatmari, and P. R. Cobbold (1993), Style and pattern of salt diapirs due to thin-skinned gravitational gliding, Campos andSantos basins, offshore Brazil, Tectonophysics, 228, 393–433.

Dewey, J. F., and J. M. Bird (1971), Origin and emplacement of the ophiolite suite: Appalachian ophiolites in Newfoundland, J. Geophys. Res.,76, 3179–3206, doi:10.1029/JB076i014p03179.

Dubois, J., C. Ravenne, A. Aubertin, J. Louis, R. Guillaume, J. Launay, and L. Montadert (1974), Continental margins near New Caledonia, in TheGeology of Continental Margins, edited by C. A. Burk and C. L. Drake, pp. 521–535, Springer, New York.

Fitzherbert, J. A., G. L. Clarke, B. Marmo, and R. Powell (2004), The origin and P-T evolution of peridotites and serpentinites of NENew Caledonia: Prograde interaction between continental margin and the mantle wedge, J. Metamorphic Geol., 22, 327–344.

Fitzherbert, J. A., G. L. Clarke, and R. Powell (2005), Preferential retrogression of high-P metasediments and the preservation of blueschist toeclogite facies metabasite during exhumation, Diahot terrane, NE New Caledonia, Lithos, 83, 67–96.

Fort, X., J. P. Brun, and F. Chauvel (2004), Salt tectonics on the Angolan margin, synsedimentary deformation processes, AAPG Bull., 88,1523–1544.

Fossen, H. (2010), Structural Geology, 463 pp., Cambridge Univ. Press, Cambridge, U. K.Ghent, E. D., J. C. Roddick, and P. M. Black (1994),

40Ar/

39Ar dating of white micas from the epidote to the omphacite zones, northern

New Caledonia: Tectonic implications, Can. J. Earth Sci., 31, 995–1001.Gonord, H. (1977), Recherches sur la géologie de la Nouvelle-Calédonie; sa place dans l’ensemble structural du Pacifique Sud-Ouest, Thèse

Doct. Etat, Univ. Montpellier,France, 341 p.Gonord, H., J. L. Saos, and J. Leguéré (1973), Note préliminaire sur la présence de décrochements en Nouvelle-Calédonie centrale, C. R. Acad.

Sci. Paris, 277, 1841–1844.Guillon, J. H. (1975), Les massifs péridotitiques de Nouvelle-Calédonie, Mémoires, vol. 76, 120 pp., ORSTOM, Paris.Guillon, J. H., and P. Routhier (1971), Les stades d’évolution et de mise en place des massifs ultramafiques de Nouvelle-Calédonie, Bull. Bur. Rech.

Géol. Min., section IV, vol. 2, pp. 5–38, Editions du BRGM, Orléans, France.Guillot, S., S. Schwartz, B. Reynard, P. Agard, and C. Prigent (2015), Tectonic significance of serpentinites, Tectonophysics, 646, 1–19.Klingelhoefer, F., Y. Lafoy, J. Collot, E. Cosquer, L. Géli, H. Nouzé, and R. Vially (2007), Crustal structure of the basin and ridge system west of

New Caledonia (southwest Pacific) from wide-angle and reflection seismic data, J. Geophys. Res., 112, B11102, doi:10.1029/2007JB005093.Lafoy, Y. et al. (2004), Rapport de mission de la campagne ZoNéCo 11 de sismique lourde (multitraces, réfraction, haute résolution) à bord du

N/O l’Atalante (8 Sept. – 5 Oct. 2004), Ifremer, Paris, 147 p.Lafoy, Y., I. Brodien, R. Vially, and N. F. Exon (2005), Structure of the basin and ridge system west of New Caledonia (southwest Pacific): A

synthesis, Mar. Geophys. Res., 26, 37–50.Lagabrielle, Y., and A. Chauvet (2008), The role of extensional tectonics in shaping Cenozoic New Caledonia, Bull. Soc. Géol. France, 179,

315–329.Lagabrielle, Y., A. Chauvet, M. Ulrich, and S. Guillot (2013), Passive obduction and gravity-driven emplacement of large ophiolitic sheets: The

New Caledonia ophiolite (SW Pacific) as a case study?, Bull. Soc. Géol. France, 184, 545–556.Lahondère, D., S. Lesimple, F. Cagnard, A. Lahfid, G. Wille, and P. Maurizot (2012), Serpentinisation et fibrogenèse dans les massifs de

péridotite de Nouvelle-Calédonie, Bur. Rech. Géol. Min., France, Public Report BRGM/RP-60192-FR, 458 p.Leblanc, M. (1995), Chromitite and ultramafic rock compositional zoning through a paleotransform fault, Poum, New Caledonia, Econ. Geol.,

90, 2028–2039.Leguéré, J. (1976), Des corrélations entre la tectonique cassante et l’altération supergène des péridotites de Nouvelle-Calédonie, PhD thesis,

Univ. Montpellier, France, 94 p.Leveridge, B. E., and R. K. Shail (2011), The Gramscatho Basin, south Cornwall, UK: Devonian active margin successions, Proc. Geol. Assoc., 122,

568–615.Lillie, A. R. (1970), The structural geology of lawsonite and glaucophane schists of the Ouégoa district, New Caledonia, New Zealand J. Geol.

Geophys., 13, 72–116.Lillie, A. R. (1975), Structures in the lawsonite-glaucophane schists of New Caledonia, Geol. Mag., 112, 225–240.Loncke, L., V. Gaullier, J. Mascle, B. Vendeville, and L. Camera (2006), The Nile deep-sea fan: An example of interacting sedimentation, salt

tectonics, and inherited subsalt paleotopographic features, Mar. Petrol. Geol., 23, 297–315.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 24

Page 32: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Maitrepierre, L. (2012), Planches 10 et 11: Les types de temps et les cyclones, les éléments du climat, in Atlas de la Nouvelle-Calédonie,pp. 53–60, IRD-Congrès de la Nouvelle-Calédonie, Marseille-Nouméa.

Mann, P., and M. B. Gordon (1996), Tectonic uplift and exhumation of blueschist belts along transpressional strike-slip fault zones, inSubduction: Top to Bottom, Geophys. Monogr., vol. 96, edited by E. Bebout et al., pp. 143–154, AGU, Washington, D. C.

Marchesi, C., C. J. Garrido, M. Godard, F. Belley, and E. Ferré (2009), Migration and accumulation of ultra-depleted subduction-related melts inthe Massif du Sud ophiolite (New Caledonia), Chem. Geol., 266, 180–195.

Maurizot, P. (2007), Cartographie d’aide à l’aménagement dans la zone Voh-Koné-Pouembout-Poya. Phase 1: Mise à jour de la cartegéologique sur les communes de Koné et Pouembout et bilan des données existantes, Bur. Rech. Géol. Min., France, Public ReportBRGM/RP-54898-FR, 76 p.

Maurizot, P. (2011), First sedimentary record of the pre-obduction convergence in New Caledonia: Formation of an Early Eocene accretionarycomplex in the north of Grande Terre and emplacement of the “Montagnes Blanches” nappe, Bull. Soc. Géol. France, 182, 479–491.

Maurizot, P., and D. Cluzel (2014), Pre-obduction records of Eocene foreland basins in central New Caledonia: An appraisal from surfacegeology and Cadart-1 borehole data, New Zealand J. Geol. Geophys., 57, 300–311.

Maurizot, P., and M. Vendé-Leclerc (2009), Carte géologique de la Nouvelle-Calédonie à l’échelle 1/500000, Dir. de l’Industrie, des Mines et del’Energie – Service de Géologie de Nouvelle-Calédonie, Bur. Rech. Géol. Min., France.

Maurizot, P., J. P. Paris, and D. Feignier (1985a), Paléogéographie de part et d’autre de l’Accident ouest calédonien durant la période Crétacésupérieur-Paléocène: Autochtonie de la formation des basaltes de la Côte ouest, Géol. France, 1, 53–60.

Maurizot, P., et al. (1985b), Données nouvelles sur les “fils de serpentinite” de Nouvelle-Calédonie, Géol. France, 1, 61–67.Maurizot, P., J. M. Eberlé, C. Habault, and C. Tessarolo (1989), Carte géologique de la Nouvelle-Calédonie à l’échelle 1/50000, feuille

Pam-Ouégoa, 2e édition, map sheet and explanatory notes, Bur. Rech. Géol. Min., France.Maurizot, P., Y. Lafoy, and M. Poupée (2002), Cartographie des formations superficielles et des aléas mouvements de terrain en Nouvelle-

Calédonie, zone du Koniambo, Bur. Rech. Géol. Min., France, Public Report BRGM/RP51624-FR, 45 p.Merle, O. (1986), Patterns of stretch trajectories and strain rates within spreading-gliding nappes, Tectonophysics, 124, 211–222.Morley, C. K., and G. Guerin (1996), Comparison of gravity-driven deformation styles and behavior associated with mobile shales and salt,

Tectonics, 15, 1154–1170, doi:10.1029/96TC01416.Mourgues, R., E. Lecomte, B. Vendeville, and S. Raillard (2009), An experimental investigation of gravity-driven shale tectonics in prograda-

tional delta, Tectonophysics, 474, 643–656.Nicolas, A. (1989), Structures of Ophiolites and Dynamics of Oceanic Lithosphere, 367 pp., Kluwer Acad., London.Orloff, O. (1968), Etude géologique et géomorphologique des massifs d’ultrabasites compris entre Houailou et Canala (Nouvelle-Calédonie),

PhD thesis, Univ. Montpellier, France, 189 p.Paquette, J. L., and D. Cluzel (2007), U-Pb zircon dating of post-obduction volcanic-arc granitoids and a granulite-facies xenolith from

New Caledonia; inference on southwest Pacific geodynamic models, Int. J. Earth Sci., 96, 613–622.Paris, J. P. (1981), Géologie de la Nouvelle-Calédonie, un essai de synthèse, Mém. Bur. Rech. Géol. Min., vol. 113, 274 pp, Editions du BRGM,

Orléans, France.Passchier, C. W., and R. A. J. Trouw (1996), Microtectonics, 289 pp., Springer, Heidelberg, Berlin.Pirard, C., J. Hermann, and H. S. C. O’Neill (2013), Petrology and geochemistry of the crust-mantle boundary in a nascent arc, Massif du Sud

ophiolite, New Caledonia, SW Pacific, J. Petrol., 54, 1759–1792.Poutchovsky, A., and J. Récy (1982), Apports de la géologie terrestre et données nouvelles de microtectonique, in Contribution à l’étude

géodynamique du Sud-Ouest Pacifique, Travaux et Documents de l’ORSTOM (Paris), vol. 147, pp. 485–503, ORSTOM, Paris.Prinzhofer, A., A. Nicolas, D. Cassard, J. Moutte, M. Leblanc, J. P. Paris, and M. Rabinovitch (1980), Structures in the New Caledonia peridotites-

gabbros: Implications for oceanic mantle and crust, Tectonophysics, 69, 85–112.Quesnel, B., P. Gautier, P. Boulvais, M. Cathelineau, P. Maurizot, D. Cluzel, M. Ulrich, S. Guillot, S. Lesimple, and C. Couteau (2013), Syn-tectonic,

meteoric water-derived carbonation of the New Caledonia peridotite nappe, Geology, 41, 1063–1066.Quesnel, B., P. Gautier, M. Cathelineau, P. Boulvais, C. Couteau, and M. Drouillet (2016a), The internal deformation of the Peridotite Nappe of

New Caledonia: A structural study of serpentine-bearing faults and shear zones in the Koniambo Massif, J. Struct. Geol., 85, 51–67.Quesnel, B., P. Boulvais, P. Gautier, M. Cathelineau, C. M. John, M. Dierick, P. Agrinier, and M. Drouillet (2016b), Paired stable isotopes (O, C)

and clumped isotope thermometry of magnesite and silica veins in the New Caledonia Peridotite Nappe, Geochim. Cosmochim. Acta, 183,234–249.

Ramsay, J. G., and M. I. Hubert (1987), The Techniques of Modern Structural Geology, Volume 2: Folds and Fractures, 392 pp., Academic Press,London.

Rawling, T. J., and G. S. Lister (1999), Oscillating modes of orogeny in the southwest Pacific and the tectonic evolution of New Caledonia,Geol. Soc. London Spec. Publ., 154, 109–127.

Rigolot, P. (1989), Evolution morphologique et structurale de la marge occidentale de la ride de Nouvelle-Calédonie (SW Pacifique),Bull. Centres Rech. Explor-Prod. Elf-Aquitaine, 13, 319–344.

Spandler, C., D. Rubatto, and J. Hermann (2005), Late Cretaceous-Tertiary tectonics of the southwest Pacific: Insights from U-Pb sensitive,high-resolution ion microprobe (SHRIMP) dating of eclogite facies rocks from New Caledonia, Tectonics, 24, TC3003, doi:10.1029/2004TC001709.

Stern, R. J., M. Reagan, O. Ishizuka, Y. Ohara, and S. Whattam (2012), To understand subduction initiation, study forearc crust: To understandforearc crust, study ophiolites, Lithosphere, 4, 469–483.

Tissot, B., and A. Noesmoen (1958), Les bassins de Nouméa et de Bourail (Nouvelle-Calédonie), Rev. Inst. Fr. Pétrole, 13, 739–760.Titus, S. J., S. M. Maes, B. Benford, E. C. Ferré, and B. Tikoff (2011), Fabric development in the mantle section of a paleo-transform fault and its

effect on ophiolite obduction, New Caledonia, Lithosphere, 3, 221–244.Ulrich, M. (2010), Péridotites et serpentinites du complexe ophiolitique de la Nouvelle-Calédonie, PhD thesis, Univ. Nouvelle-Calédonie et

Univ. Grenoble, France, 246 p.Ulrich, M., C. Picard, S. Guillot, C. Chauvel, D. Cluzel, and S. Meffre (2010), Multiple melting stages and refertilization as indicators for ridge to

subduction formation: The New Caledonia ophiolite, Lithos, 115, 223–236.Vitale Brovarone, A., and P. Agard (2013), True metamorphic isograds or tectonically sliced metamorphic sequence? New high-spatial

resolution petrological data for the New Caledonia case study, Contrib. Mineral. Petrol., 166, 451–469.Willett, S. D. (1999), Orogeny and orography: The effects of erosion on the structure of mountain belts, J. Geophys. Res., 104, 28,957–28,981,

doi:10.1029/1999JB900248.

Tectonics 10.1002/2016TC004318

GAUTIER ET AL. OBDUCTION IN NEW CALEDONIA 25

Page 33: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Geology and deformation of the ophiolitic

nappe in New Caledonia

Quesnel, Benoit, Gautier, Pierre, Cathelineau, Michel BoulvaisPhilippe, Couteau Clément, et Drouillet Maxime. (2016)

The internal deformation of the Peridotite Nappe of New Caledonia: A structural study of serpentine-bearing faults and

shear zones in the Koniambo Massif

Journ. Struct. Geol. : 85, 51-67 .

- 2 -

Page 34: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

The internal deformation of the Peridotite Nappe of New Caledonia: Astructural study of serpentine-bearing faults and shear zones in theKoniambo Massif

Benoît Quesnel a, *, Pierre Gautier a, Michel Cathelineau b, Philippe Boulvais a,Cl�ement Couteau c, Maxime Drouillet c

a Universit�e Rennes 1, CNRS, G�eosciences Rennes UMR 6118, OSUR, 35042 Rennes Cedex, Franceb Universit�e de Lorraine, CNRS, CREGU, GeoRessources Lab., 54506 Vandoeuvre-l�es-Nancy, Francec Service G�eologique, Koniambo Nickel SAS, 98883 Voh, New Caledonia

a r t i c l e i n f o

Article history:Received 11 January 2016Received in revised form17 February 2016Accepted 20 February 2016Available online 23 February 2016

Keywords:New CaledoniaOphioliteSerpentineMagnesiteNappeStructural analysis

a b s t r a c t

We present a structural analysis of serpentine-bearing faults and shear zones in the Koniambo Massif,one of the klippes of the Peridotite Nappe of New Caledonia. Three structural levels are recognized. Theupper level is characterized by a dense network of fractures. Antigorite and polygonal serpentine formslickenfibers along fault planes with distinct kinematics. As a result, the upper level keeps the record of atleast two deformation events, the first associated with the growth of antigorite (WNW-ESE extension),the second with the growth of polygonal serpentine (NWeSE compression). The lower level coincideswith the ‘serpentine sole’ of the nappe, which consists of massive tectonic breccias overlying a layer ofmylonitic serpentinites. The sole records pervasive tangential shear with top-to-SW kinematics andrepresents a d�ecollement at the base of the nappe. The intermediate level is characterized by thepresence of several meters-thick conjugate shear zones accommodating NEeSW shortening. Like thesole, these shear zones involve polygonal serpentine and magnesite as the main syn-kinematic mineralphases. The shear zones likely root into the basal d�ecollement, either along its roof or, occasionally,around its base. Compared to top-to-SW shearing along the sole, the two deformation events recorded inthe upper level are older.

The three structural levels correlate well with previously recognized spatial variations in the degree ofserpentinization. It is therefore tempting to consider that the intensity of serpentinization played a majorrole in the way deformation has been distributed across the Peridotite Nappe. However, even the leastaltered peridotites, in the upper level, contain so much serpentine that, according to theoretical andexperimental work, they should be nearly as weak as pure serpentinite. Hence, no strong verticalgradient in strength due to variations in the degree of serpentinization is expected within the exposedpart of the nappe. Our proposal is that strain localization along the serpentine sole results from thejuxtaposition of the nappe, made of weak serpentinized peridotites, against the strong mafic rocks of itssubstratum. This interpretation is at odds with the intuitive view that would consider the nappe, made ofperidotites, as stronger than its basement.

© 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Ophiolites commonly occur within orogenic belts as a result ofthe obduction of a piece of oceanic lithosphere upon a continentalbasement (Coleman, 1971; Dewey and Bird, 1971). The ultramaficrocks that constitute a large part of most ophiolitic nappes gener-ally display evidence for pronounced fluid-rock interactions in theform of a spatially uneven development of serpentinization. Due to

* Corresponding author.E-mail addresses: [email protected] (B. Quesnel), pierre.gautier@univ-

rennes1.fr (P. Gautier), [email protected] (M. Cathelineau),[email protected] (P. Boulvais), [email protected](M. Drouillet).

Contents lists available at ScienceDirect

Journal of Structural Geology

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

http://dx.doi.org/10.1016/j.jsg.2016.02.0060191-8141/© 2016 Elsevier Ltd. All rights reserved.

Journal of Structural Geology 85 (2016) 51e67

Page 35: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

their low strength in comparison to peridotites (e.g., Escartín et al.,2001), serpentinites have the capacity to promote strain localiza-tion and greatly influence the style of deformation on the litho-sphere scale (e.g., Hirth and Guillot, 2013). Several types ofserpentine may be identified in a single ophiolite, which raises thequestion of their relative chronology and their geodynamic envi-ronment of formation (e.g., Coulton et al., 1995). The topic of thiswork is on ophiolitic nappes remaining near the surface, thusexcluding the case of dismembered meta-ophiolitic sequences oc-casionally exposed in the internal zones of orogenic belts. In prin-ciple, serpentinization within an ophiolitic nappe may witness thesuccessive stages of oceanic accretion, intra-oceanic subduction(the ophiolite lying in the upper plate, for instance in the fore-arcregion), obduction, and post-obduction evolution. Post-obductionevents may involve post-convergence extension and/or collisionalorogeny.

The Peridotite Nappe of New Caledonia is one of the fewophiolites in the world to have escaped collision after obduction(e.g., Cluzel et al., 2012). In this case, deciphering early serpenti-nization events should be easier. Recent work has shown that thePeridotite Nappe hosts various types of serpentine (Lahond�ere andMaurizot, 2009; Ulrich, 2010; Lahond�ere et al., 2012). An intimateassociation between deformation and the occurrence of serpen-tines is recognized in these studies, however no attempt has beenmade so far to characterize this deformation. More generally,structural studies carried out on the ultramafic section of ophioliticnappes usually focus on serpentine-free high grade fabrics (e.g.,Boudier et al., 1988; Suhr and Cawood, 1993; Cook et al., 2000).Studies including an analysis of serpentine-bearing faults andshear zones are relatively rare (Bailey et al., 2000; Titus et al., 2002;Schemmann et al., 2007; La�o-D�avila and Anderson, 2009; Federicoet al., 2014).

In New Caledonia, earlier studies concerned with the PeridotiteNappe have focused on i) the high temperature intra-oceanicstages of deformation within the nappe (Prinzhofer et al., 1980;Titus et al., 2011), ii) the obduction process, however basedessentially on the analysis of the rock units underlying the nappe(e.g., Cluzel et al., 2001, 2012; Spandler et al., 2005; Lagabrielleet al., 2013) and iii) post-obduction brittle extension that shapedthe island as a large horst (e.g., Lagabrielle et al., 2005; Chardonand Chevillotte, 2006). Furthermore, much attention has beenput on the chemistry and mineralogy of thick lateritic profilesformed at the expense of the peridotites, which led to the devel-opment of worldclass nickel mineralizations (e.g., Orloff, 1968;Trescases, 1975; Butt and Cluzel, 2013). In contrast, only limitedwork has concerned serpentines-bearing faults and shear zonesthat are ubiquitous throughout the Peridotite Nappe, despite thecommon statement that the distribution of nickel ore is highlydependent on this fault network (Legu�er�e, 1976; Cluzel and Vigier,2008).

In this study, we had access to a recently opened large miningsite in the Koniambo Massif, one of the klippes of the PeridotiteNappe in northwestern New Caledonia (Fig. 1). This site provides aquasi-continuous exposure of fresh rocks from the base of thenappe, near sea level, up to ~800 m elevation, where the mainlaterite capping the peridotites occurs. We focused on the study ofstructures and deformation associated with serpentines. Thesedata enable to describe the internal deformation of the nappe inrelation with successive serpentinization events. The implicationsof this study for the tectonic evolution of New Caledonia are notdiscussed here and will be the topic of a distinct paper.

2. Geological setting

New Caledonia is located in the southwest Pacific Ocean,

1300 km east of Australia (Fig. 1). On the main island, known as the“Grande Terre”, the Peridotite Nappe overlies, with a sub-horizontal tectonic contact (Avias, 1967; Guillon, 1975), a substra-tum composed of several volcano-sedimentary units (Paris, 1981;Cluzel et al., 2001, 2012). The Peridotite Nappe is essentiallyexposed in the “Massif du Sud”, covering much of the southeasternthird of the Grande Terre, and as a series of klippes along thenorthwestern coast (Fig. 1). In the Massif du Sud, the thickness ofthe nappe is at least 1.5 km and may reach 3.5 km (Guillon, 1975).The nappe is mostly composed of harzburgites except in thenorthernmost klippes where lherzolites dominate (e.g., Ulrichet al., 2010). Compositional layering is essentially represented by1e100 m-thick layers of dunite within the harzburgites. In themain part of the nappe, the degree of serpentinization of the pe-ridotites is variable but moderate (Orloff, 1968), serpentinesoccurring preferentially along a network of mm-to ~10 cm-thickfractures and shear zones (e.g., Legu�er�e, 1976; Lahond�ere andMaurizot, 2009; Lahond�ere et al., 2012). Fracture infillings andveins with various textures are described, containinglizardite ± chrysotile or antigorite ± chrysotile and/or an additionaltype of light-colored serpentine not firmly identified (Lahond�ereand Maurizot, 2009; Lahond�ere et al., 2012). Serpentinization be-comes pervasive along the base of the nappe, forming a ‘sole’ inwhich deformation has been intense (Avias, 1967; Orloff, 1968;Guillon, 1975; Legu�er�e, 1976; Cluzel et al., 2012; Quesnel et al.,2013). The thickness of this sole is a few tens of meters in theMassif du Sud but reaches a few hundred meters in the north-western klippes (Guillon, 1975; Maurizot et al., 2002). According toUlrich (2010), serpentinization of the sole involved three stages,firstly a pervasive development of lizardite, secondly the formationof antigorite within mm to cm-thick veins, finally the formation ofchrysotile in local veinlet networks. In some of the veinlets, Ulrich(2010) also reports the partial replacement of chrysotile bypolygonal serpentine. Recently, Ulrich et al. (2014) established thatpolygonal serpentine is more widespread in the sole than previ-ously thought, forming light green serpentine veins associatedwith chrysotile veinlets.

Following their exposure to aerial conditions, the peridotiteshave been subject to intense weathering under dominantly warmand wet climatic conditions. This has led to the development oflaterites up to ~30m thick (e.g., Chevillotte et al., 2006). Leaching ofthe peridotites by downward infiltrating meteoric waters has

Fig. 1. Simplified geological map of the Grande Terre, New Caledonia, showing theexposures of ultramafic rocks. Most of these rocks are originally part of the PeridotiteNappe. Adapted from Maurizot and Vend�e-Leclerc (2009).

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6752

Page 36: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

resulted in a redistribution of elements, leading to high nickelconcentrations at the base of the weathering profile, in the tran-sition zone between coarse-grained saprolite and the bedrock(Orloff, 1968; Trescases, 1975; Legu�er�e, 1976; Paris, 1981). Someveins with nickel-rich mineralizations show evidence for a syn-tectonic emplacement (Cluzel and Vigier, 2008). Magnesite veins,which frequently occur along the serpentine sole of the PeridotiteNappe, represent another by-product of the leaching of the peri-dotites by meteoric fluids (Ulrich, 2010; Quesnel et al., 2013).Quesnel et al. (2013) showed that at least some of these veins havebeen emplaced during pervasive shearing along the sole.

In the KoniamboMassif, the Peridotite Nappe is exposed from itsbase, near sea level, up to ~800 m elevation (Fig. 2a,b). Map re-lations (Carrou�e, 1972; Maurizot et al., 2002) show that the basalcontact is subhorizontal aroundmuch of the massif but delineates alarge open antiform in the west (Fig. 2b). At elevations above~400 m, the nappe is capped by a highly dissected and partlyreworked lateritic profile (Maurizot et al., 2002). At lower eleva-tions, laterites of the westerly-inclined Kaf�eat�e plateau probablybelong to a younger planation surface (Latham, 1977; Chevillotteet al., 2006). The Koniambo Massif essentially consists of harzbur-gites with interlayers of dunite that define a crude compositionallayering with a fairly regular ENE-WSW strike and a ~50� south-ward dip (Maurizot et al., 2002). With regard to the degree of ser-pentinization, Maurizot et al. (2002) distinguished three main rocktypes, namely (i) moderately serpentinized peridotites inwhich theprimary compositional layering is preserved, (ii) highly serpenti-nized peridotites and (iii) massive serpentinites. The latter form theserpentine sole of the nappe, about 200 m thick. The highly ser-pentinized peridotites overlie the sole and form a distinct layerwhich, according to Maurizot et al. (2002), is ~200 m thick on the

western flank of the massif but thins out further east (Fig. 2a,b). Themoderately serpentinized peridotites occupy the higher part of themassif.

3. Methods

3.1. Strategy

Field work focused on fresh rock exposures provided by the last~5 years of mining activity in the Koniambo Massif. The serpentinesole is well exposed in the Vavouto peninsula, especially along twocross-sections here named VA1 (700m long, striking N082�, Fig. 2c)and VA2 (a composite section ~330m longwith amean NeS strike).The VA2 section shows the basal contact of the nappe and itssubstratummade of basalts and minor cherts (Fig. 3a). Further east,the mine access road runs from the base of the nappe (height spot65 m in Fig. 2a) to the upper part of the massif; it provides goodoutcrops of the intermediate level of highly serpentinized perido-tites. The moderately serpentinized peridotites are exposed athigher levels of the access road and in a series of currently exploitedopen pits. We carried out preliminary field observations andsampled rocks with various textures and fabrics (veins, fractureinfillings, fault coatings, cleavage domains within shear zones). Wedetermined the mineralogical content of the samples, which thenserved as references for macroscopic assessment of the mineralogywithin specific structures during renewed field work and structuralanalysis.

3.2. Analytical methods

Reference samples have been characterized through a multi-

Fig. 2. a) Geological map of the central and southern parts of the Koniambo Massif, adapted from Carrou�e (1972) and Maurizot et al. (2002). b) General cross-section of theKoniambo Massif, located in (a). c) Report of the main fault zones along the VA1 section, in the Vavouto peninsula.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 53

Page 37: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

technique approach including transmission electron microscopy,microprobe analysis (not presented here) and Raman spectroscopy.Raman spectroscopy measurements were carried out at the labo-ratory G�eoRessources Nancy, France, using a Horiba Jobin-YvonLabram HR800 spectrometer and a visible ionized argon lasersource with a wavelength of 514 nm. The output power of the laserwas 100 mW and measurements were performed using anOlympus lens of x50 to focus the laser beam onto an area 1 mm indiameter. Spectra are the average of 6e10 acquisitions of 20 s eachto optimize the signal/noise ratio. For serpentine identification,only one region of the Raman spectrum has been investigated,between 3520 cm�1 and 3870 cm�1 (Fig. 4). This region enables tocharacterize the hydroxyl groups which are the most discriminantfor distinguishing the polymorphs of serpentine (Auzende et al.,2004; Schwartz et al., 2013; Ulrich et al., 2014).

For analyzing the deformation, we focused our attention onshear zones and fault planes. The kinematics of shear zones havebeen determined using the obliquity of cleavagewith respect to theshear zone walls (Fig. 5e,f) and, occasionally, the occurrence of a setof oblique shear bands. For determining the kinematics of faults,the offset of pre-kinematic markers or the presence of subsidiaryRiedel shears has occasionally been used, however inmost cases wefavored the use of accreted slickenfibers with a clear staircase ge-ometry (Fig. 5a,b,c,d). The advantage of stepped slickenfibers is toprovide a robust sense-of-slip indicator and to allow establishingwhich mineral or mineral assemblage was stable during a givenfault slip event (Durney and Ramsay, 1973; see also, e.g., Twiss andMoores, 1992).

The bulk deformation of densely faulted areas has been assessedusing the FaultKin software of Allmendinger et al. (2012). The basictenets of this software, described by Marrett and Allmendinger(1990), are the same as those of the right-dihedra method(Angelier and Mechler, 1977; Angelier, 1994). It should be noticedthat the approach used by Marrett and Allmendinger (1990) isintended to provide a qualitative description of the strain ellipsoid,not the stress ellipsoid. Several authors have argued that, inessence, the analysis of fault slip data informs more about strain

than about stress (e.g., Twiss and Unruh, 1998; Tikoff and Wojtal,1999; Gapais et al., 2000). According to C�el�erier et al. (2012),methods aiming at retrieving paleostresses may be more suited forseparating distinct subsets (presumed to witness successive tec-tonic events) within a given fault slip dataset whereas methodsaiming at retrieving strain (e.g., Marrett and Allmendinger, 1990)may be more suited for yielding an average estimate of strainwithin a rock volume. In our case, the mineralogy of slickenfibers,which is of two types, has provided an independent criterion forseparating the fault slip data into two subsets, therefore the ‘strain’approach seems more appropriate.

4. Results

4.1. Reference samples

We here describe three hand specimens that are representativeof field cases where serpentine occurs in association with defor-mation. The two first specimens come from higher levels of theKoniambo Massif. The third sample comes from the Vavoutopeninsula, within the serpentine sole.

The first sample includes a ~1 cm-thick vein of light olive greenserpentine with a massive aspect (Fig. 4a). The vein has been activeas a fault plane as indicated by fine striations underlined by blackoxides along a surface parallel to its margin. The Raman spectrumof the veinmaterial (analysis #1e1) shows two closely spaced mainbands centered on ~3688 cm�1 and ~3694 cm�1, which are diag-nostic of polygonal serpentine (Auzende et al., 2004; Ulrich, 2010;Ulrich et al., 2014).

The second sample includes a ~1 cm-thick vein of platy-fibrousserpentine (Fig. 4b). Outside the vein, the host rock shows a Ramanspectrum (#2e1) with two main bands centered on ~3688 cm�1

and ~3696 cm�1, indicating the presence of polygonal serpentine.The vein itself is composite. Two insulated domains with diffuseboundaries are made of darker serpentine showing a Ramanspectrum (#2e2) with two main bands centered on ~3672 cm�1

and ~3700 cm�1, typical for antigorite. The main part of the vein

Fig. 3. Two field views at the (a) northern and (b) southern ends of the VA2 section, in the Vavouto peninsula (location in Fig. 2a).

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6754

Page 38: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

appears as an heterogeneous stack of fibers. The Raman spectrumof analysis #2e3 in this domain shows a minor band centered on~3671 cm�1, attesting for the presence of antigorite. In addition, themain band centered on ~3698 cm�1 and the adjacent poorlydeveloped band at ~3688 cm�1 indicate the presence of polygonalserpentine or/and chrysotile. Finally, the margin of the vein isunderlined by small domains of homogeneous light greenserpentine. Their staircase distribution indicates that they formedas the result of a fault slip event. The Raman spectrum of analysis#2e4 shows a main band centered on ~3697 cm�1 and an adjacentpoorly developed band at ~3690 cm�1 which, again, points to thepresence of polygonal serpentine or/and chrysotile. In that case,however, chrysotile is unlikely because the serpentine in this partof the vein is massive and does not display the habitus of welldefined submillimetric fibers that is typical for chrysotile(Lahond�ere and Maurizot, 2009; Ulrich, 2010; Lahond�ere et al.,2012). The small domains of polygonal serpentine have welldefined boundaries, which suggests that they formed later than theadjacent domain of darker antigorite did. Furthermore, the platy-fibrous nature of the main part of the vein is typical for antigorite(e.g., Lahond�ere et al., 2012), therefore the occasional intimate as-sociation of antigorite and polygonal serpentine, like in analysis#2e3, can be interpreted as the result of a partial pseudomorphosis

of antigorite by fine-grained polygonal serpentine. Thus, the crys-tallization of polygonal serpentine appears to postdate that ofantigorite.

The third sample consists of a fist-sized elongate clast takenfrom one of the major shear zones met along the VA1 section(Fig. 2c). The clast, which is made of pervasively serpentinizedharzburgite with few orthopyroxene relics, is embedded in a filmup to 0.5 cm thick of homogeneous pale green serpentine(Fig. 4c,d). The Raman spectrum (analysis #3e1) shows two closelyspaced main bands centered on ~3691 cm�1 and ~3697 cm�1,indicating that the film consists of polygonal serpentine.

4.2. Field determination of the serpentine polymorphs associatedwith deformation

Based on the characterization of the three above samples as wellas other reference samples, we have established a number ofmacroscopic criteria for a field assessment of the mineralogy offault zones and shear zones. Within fault zones, antigorite andpolygonal serpentine are both able to build macroscopic slick-enfibers with a clear staircase geometry. For distinguishing the twopolymorphs, the main argument is textural. Antigorite has a wellexpressed platy-fibrous aspect (see also, e.g., Lahond�ere et al., 2012)

Fig. 4. Close views of the (a) first, (b) second and (c & d) third reference samples discussed in the text, and corresponding Raman spectra.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 55

Page 39: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

with fibers frequently several centimeters long (Fig. 5a). In contrast,polygonal serpentine is massive, matte, and develops only shortslickenfibers (Fig. 5b,c). A second argument is the fact that polyg-onal serpentine is pale greenwhereas antigorite is generally darker(Figs. 4b,c,d and 5a,b,c). However, in the saprolitic levels, weath-ering modifies the colors and polygonal serpentine tends tobecome olive green (Figs. 4a and 5c). In addition, some fault planesbear slickenfibers showing mixed properties, with a clear platy-fibrous nature but a pale green color and a matte aspect (Fig. 5d).In that case, taking into account the likeliness of a pseudomorphicreplacement of antigorite by polygonal serpentine, as in the secondreference sample above, we have assumed that the slickenfiberswitness a syn-antigorite slip event.

Antigorite and polygonal serpentine are also involved in shearzones. There, antigorite forms distinct lamellae (Fig. 5e) whereaspolygonal serpentine is more massive despite the development of apronounced cleavage (Fig. 5f). Most of the shear zones we haveobserved involve polygonal serpentine. There is always a closespatial coincidence between the shear zone walls and the marginsof the serpentine body (Fig. 5e,f). The fabric in each shear zone ispronounced and, in principle, shearing might post-date serpenti-nization. Nevertheless, would shearing occur under physical con-ditions outside the stability range of a primary serpentine, it isexpected that a new serpentine polymorph would easily crystallizeduring the pervasive deformation. Hence, for shear zones with anhomogeneous serpentine content, we assume that shearing andmineral growth were coeval.

Lizardite and chrysotile were also observed in the field, withfeatures similar to those reported by Ulrich (2010) and Lahond�ere

et al. (2012). Dark lizardite is widespread, present as a diffusegrain-scale network in the peridotites and as infillings of~1 mme~5 cm-wide joints. It is also found on the margins of manyof the faults and shear zones that involve antigorite or polygonalserpentine (Fig. 5c,e). However, lizardite itself only rarely bearsstriations and, in our experience, never builds macroscopic slick-enfibers, therefore we made no attempt to characterize the defor-mation associated with lizardite. Chrysotile occurs as very thinfibers always oriented at right angle to the veinwalls within�1 cm-thick veinlets. Dense networks of subparallel veinlets are locallyobserved. Such veins can be interpreted as tension gashes, there-fore, in principle, they could provide some clue on the tectonicregime that prevailed during chrysotile growth. Alternatively,tension gashes may open in the absence of a significant tectoniccontribution, for instance due to fracturing assisted by fluid pres-sure. For this reason, and because these thin veins represent verysmall amounts of strain, we ignored them during the analysis ofdeformation.

4.3. Structural analysis

We identified three structural levels, which correlate with therock horizons identified byMaurizot et al. (2002) on the basis of theintensity of serpentinization (Fig. 2a,b).

4.3.1. Upper structural levelThe Upper structural level coincides with the domain of

moderately serpentinized peridotites at higher levels of the massif.A dense network of fractures, most of them dipping steeply,

Fig. 5. Field views of serpentine-bearing (a to d) fault planes and (e & f) shear zones in the Upper structural level.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6756

Page 40: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

crosscut the peridotites (Fig. 6a). The spacing between adjacentmacroscopic fractures is always less than ~2 m. In between frac-tures, the least serpentinized peridotites show ‘mesh’ textures (e.g.,Roum�ejon and Cannat, 2014) resulting from a partial replacementof olivine by lizardite along grain boundaries and grain-scalemicrofractures (Fig. 6b). On the outcrop scale, in addition to ‘dry’joints, many fractures are filled with one or several types ofserpentine. When present together with another serpentine poly-morph, lizardite occupies the margins of the fracture (Fig. 5c).When occurring together, antigorite and polygonal serpentineoccupy the core vs. the margins of the fracture, respectively. Asdiscussed with the second reference sample in Section 4.1 (Fig. 4b),polygonal serpentine formed later.

In most veins involving fibrous polygonal serpentine or/andantigorite, the fibers lie at a very low angle to the vein walls, hencethese veins can be interpreted as fault zones (e.g., Durney andRamsay, 1973). A total of 125 such faults has been measured in aseries of open pits covering a relatively small area on the top of theKoniambo Massif (Fig. 2a). The syn-polygonal serpentine vs. syn-antigorite nature of each fault displacement has been determinedby assessing the mineralogy of associated slickenfibers(Fig. 5a,b,c,d) (see Section 4.2). As a result, two fault subsets havebeen defined, which compare well in terms of orientation but havedistinct kinematics (Fig. 7a,b). The majority of the faults withdominantly dip-slip movement have a normal sense of slip whenassociated with antigorite vs. a reverse sense of slip when associ-ated with polygonal serpentine. Fault slip analysis documents twostrikingly different strain ellipsoids (Fig. 7c,d). The syn-antigoritefaults yield a highly constrictional ellipsoid, as illustrated by thevalue of 0.17 computed for the R ratio (R¼(ε2�ε3)/(ε1�ε3)). Conse-quently, l1, the axis of maximum stretching, is the most accuratelydefined; it lies horizontally with a N102� trend (Fig. 7c). The axis ofmaximum shortening, l3, plunges steeply (64�) to the north-northeast. The syn-polygonal serpentine faults yield an ellipsoid

in the flattening field (R ¼ 0.71). Consequently, l3 is the mostaccurately defined axis; it lies subhorizontally with a N144� trend,while l1 plunges shallowly in the N057� direction (Fig. 7d). Fig. 8illustrates the results of fault analysis on the scale of a pit ~200 mwide within the main area of investigation (Pit 208) and on anisolated site located 2 km further northwest (Bilboquet, see Fig. 2afor location). Fault slip analysis on these sites yields results that fitclosely with those deduced from the bulk dataset.

Further southwest and at lower elevation with respect to theopen pits, the mine access road provides a large outcrop near thetransition between the upper and intermediate structural levels(Roadsite ‘620 m’ in Fig. 2a). It shows fault planes with slickenfibersas well as shear zones a few centimeters thick (Fig. 5f), bothinvolving polygonal serpentine. Fault slip analysis yields an ellip-soid in the flattening field (R ¼ 0.66), with l3 shallowly plungingsouthward (27�) along a N006� trend (Fig. 9). This is broadlyconsistent with the syn-polygonal serpentine strain field deducedfrom the open pits (Fig. 7d), though with a ~40� clockwise rotationin the direction of l3. This difference could reflect differential large-size block rotations post-dating the measured fault slip increments.However, in this area, dunitic interlayers retain the same ENE-WSWstrike and ~50� southward dip as elsewhere in the KoniamboMassif (Maurizot et al., 2002), therefore this hypothesis is unlikely.Hence, the difference is more likely to reflect a modification of thestrain field through space or time.

4.3.2. Intermediate structural levelIn the Intermediate structural level, the degree of serpentini-

zation is overall high but laterally variable. The main part is similarto exposures of the Upper structural level, with a dense network ofserpentine-bearing fractures. The distinctive feature of the Inter-mediate structural level is that this rock mass is crosscut by low-dipping shear zones a few meters thick, characterized by a perva-sive development of light green polygonal serpentine (Fig. 10a).

Fig. 6. a) Field view of an open pit with a dense network of fractures, typical from the Upper structural level (location in Fig. 2a). b) Photomicrograph of a sample representative ofthe least serpentinized peridotites within the Koniambo Massif, from the Upper structural level, in between the macroscopic fractures. The sample shows a typical ‘mesh’ texturecomposed of lizardite (Liz) mesh rims surrounding olivine (Ol) and orthopyroxene (Opx) mesh cores.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 57

Page 41: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Two such shear zones (SZ1 and SZ2) occur at high levels of theIntermediate structural level (Fig. 2a). The NEeSW horizontalspacing between the two shear zones is ~450 m. Continuous rockexposure along themine access road enables to note that no similarshear zone exists between SZ1 and SZ2. Both SZ1 and SZ2 involvepolygonal serpentine with a closely spaced cleavage as well asmagnesite (Fig. 10b,c,d,e). In the case of SZ1, striations have notbeen observed. The direction of shearing is assumed to lie at rightangle to the intersection between the cleavage and shear planeslocated along the margins of the shear zone (Fig. 10b,c), along a~N034� trend (Fig. 10f). In the case of SZ2, a major fault plane in thecore of the shear zone bears striations with a N039� trend while therelationship between cleavage and shear bands suggests a ~N085�

trend for the direction of shearing (Fig.10e,f). Both shear zones havetop-to-SW kinematics. SZ1 has the attitude of a reverse-slip shearzone (Fig. 10a,b,c,f) while SZ2 combines reverse and dextral dis-placements (Fig. 10e,f).

At lower elevations, the mine access road crosscuts a prominentSW-dipping fault zone involving a ~30 m-thick pinch of fine-

grained sediments (Maurizot et al., 2002) (Fig. 11a). These sedi-ments are originally part of either the ‘Poya terrane’ (Maurizot et al.,2002) or, more likely, the ‘Nepoui flysch’ (D. Cluzel, personalcommunication, 2012). As a result, they represent a fragment of thevolcano-sedimentary series that immediately underlie the Perido-tite Nappe (Cluzel et al., 2001). Hence, the fault zone likely rootsslightly beneath the basal contact of the nappe, or along it if onetakes into account the fact that sheets of serpentinite are frequentlymixed with substratum rocks around several klippes of the nappe(e.g., Maurizot et al., 1985). Within the fault zone, the main faultsuperposes highly sheared serpentinites onto the sediments(Fig. 11a). Its orientation is N146�, 38ºSW and it bears pronouncedstriations with a N072� trend (Fig. 11b). Along this contact andacross a structural thickness of at least 75 m above it, shear sensecriteria indicate a reverse-slip (top-to-ENE) displacement(Fig. 11c,d). This movement is consistent with the incorporation ofrocks from the substratum into the fault zone, and with the prob-able offset of the roof of the serpentine sole across it (Fig. 2a,b). Thistop-to-ENE fault can be viewed as a conjugate shear with respect to

Fig. 7. Lower hemisphere, equal-area projection showing the orientation of all measured (a) syn-antigorite and (b) syn-polygonal serpentine faults at higher levels of the KoniamboMassif (location in Fig. 2a), and (c & d) the result of the analysis of these fault slip data using the FaultKin software of Allmendinger et al. (2012). Kamb contours are shown for theaxes of maximum stretching (in c) or maximum shortening (in d) with intervals in shades of grey as depicted in the legend. The parameter R equates (ε2�ε3)/(ε1�ε3) and describesthe shape of the strain ellipsoid (R is 0 for pure constriction, 0.5 for plane strain, and 1.0 for pure flattening). Nmax is the number of faults compatible with the computed ellipsoid.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6758

Page 42: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the top-to-SW SZ1 and SZ2 shear zones.

4.3.3. Lower structural level (the serpentine sole)The best exposures of the serpentine sole lie in the Vavouto

peninsula (Figs. 2, 3,12 and 13). Further east, themap contour of thebasal contact of the nappe implies a culmination of this contactslightly northwest of the Tambounan Peak (Fig. 2a). As a result, thesurface probably dips westward beneath the peninsula, with aslope of ~6e7� (Fig. 2b). The contact is exposed at the easternmargin of the peninsula, displaying ~ EeW-trending open folds ona ~100 m scale (Fig. 3a).

The rocks along the sole are brecciated and/or foliated. Brecciasare well developed along the VA1 section (Fig. 12a) and consist ofclasts of serpentinized peridotite separated by serpentine joints(Fig. 4d). Many joints bear very fine striations (Fig. 4c). The stria-tions have extremely variable orientations, which probably resultsfrom the complexity of clast displacements/rotations during brec-ciation. Major shear zones occur along the VA1 section (Fig. 2c),marked by a reduction in clast size (Figs. 12a and 13a). The shearzone boundaries are dominantly diffuse, but locally change tosharp. The width of the main shear zones ranges from ~2 to ~10 m.In the shear zones, the greater density of fault planes subparallel totheir walls tends to produce elongate clasts that locally define acrude foliation (e.g., in the upper part of Fig. 13b). In the core ofsome shear zones, the planar fabric is more pronounced and locallyoblique to the shear zone walls. In cases where this oblique fabrichad the appearance of a closely spaced cleavage, it helped to assessthe kinematics of the shear zone. In most cases, however, it isdifficult to decidewhether the fabric could alternatively represent adense pattern of Riedel shears, hence implying the opposite sense

Fig. 8. Lower hemisphere, equal-area projection showing the orientation of faults attwo specific sites (location in Fig. 2a), and the result of the analysis of these fault slipdata.

Fig. 9. Lower hemisphere, equal-area projection showing the orientation of faults andminor shear zones at roadsite ‘620 m’ (location in Fig. 2a), and the result of the analysisof this fault slip dataset. Kamb contours are shown for the axes of maximum short-ening with intervals in shades of grey as depicted in the legend.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 59

Page 43: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

of shear compared to that deduced from an oblique cleavage. As aresult, the kinematics of individual shear zones and shear planesalong the VA1 section has essentially been deduced from their ef-fect (offset and drag folding) on pre-existing markers (Figs. 12b and13b).

The largest part of the VA2 section consists of serpentinites witha pronounced planar fabric (Figs. 3a and 12c). On a 10 cm scale, thisfabric is underlined by a disjunctive anastomosing cleavage(Fig. 13c). When present, clasts are isolated, hence the rock retains amatrix-supported texture (Fig. 12d). Generally, the clasts have anelongate and subangular to rounded shape with a consistentorientation throughout the rock (Fig. 12d). As a result, the rock hasthe typical appearance of a mylonite. Early compositional layeringhas been transposed into parallelism with the cleavage (Fig. 12c),attesting for very large strains. Under microscope, the fabric of thefoliated serpentinites is irregularly distributed but locally pervasive(e.g., in the right part of Fig. 13d). A component of ductile

deformation is thus involved. The northern end of the VA2 sectionshows that the foliated serpentinites occur immediately above thebasal contact of the nappe (Fig. 3a). The southern end of the sectionshows that they underlie the part of the sole dominated by brecciasthrough a relatively sharp transition zone (Fig. 3b). Hence, thefoliated serpentinites form a distinct layer of most intenselydeformed rocks at the base of the serpentine sole, with a thicknessof at least 20 m.

As illustrated by the third reference sample in Section 4.1(Fig. 4c,d), our observations indicate that the major shear zonesalong the VA1 section involve pale green polygonal serpentine asthe main syn-kinematic mineral phase (Figs. 12a and 13a,b). Thisalso applies to the layer of foliated serpentinites of the VA2 section(Fig. 12c,d and 13c). Magnesite also occurs within this layer andalong the VA1 section, in the form of irregularly distributed veinsand local stockworks (Figs. 2c, 12c and 13a). Magnesite occurs inclose association with polygonal serpentine (Figs. 12c,d and

Fig. 10. Field views of (a to d) SZ1 and (e) SZ2, two major low-dipping shear zones typical from the Intermediate structural level (location in Fig. 2a). Plg, polygonal serpentine; Mgs,magnesite. f) Lower hemisphere, equal-area projection showing the orientation of structural elements in the two shear zones (see the text).

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6760

Page 44: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

13a,b,c). Evidence of the syn-kinematic nature of the magnesiteveins is shown in Fig. 13 (see also Quesnel et al., 2013). In themacroscopic examples b and c, the veins are offset by local shearplanes that contribute to the bulk shearing deformation. In additionto the vein infilling habitus, elongate clusters of partly coalescentmillimetric to centimetric nodules of magnesite are occasionallyfound. Fig. 13d shows part of a pluricentimetric nodule under mi-croscope. Its right margin is irregular, made up of submillimetricnodules of cryptocrystalline magnesite. This geometry, whichsuggests an unconstrained growth of magnesite, could be taken as

an indication that magnesite has formed in the absence ofcontemporaneous deformation, after the development of the shearfabric visible in the host serpentinite. However, the uppermargin ofthe nodule is straight and coincides with one of the shear bandsrelated to this fabric. The relations between the nodule and theshear band, highlighted in the enlargement, are clearly more in linewith the shear band post-dating the nodule thanwith the opposite.Overall, the relations seen in Fig. 13d are consistent with magnesitehaving grown during shearing.

The asymmetric distribution of shear zones along the VA1(Fig. 2c) and VA2 (Fig. 12c) sections documents non coaxial defor-mation across the serpentine sole. The exposure surface beingsmooth along both sections, the orientation of the main shearzones is difficult to evaluate. Nevertheless, a dozen shear planes ofsignificant size (several meters) have been measured along the VA1section (Fig. 12e). Their mean orientation is N141�, 46� SW. Thissuggests top-to-SW kinematics for the bulk shearing deformation.Top-to-SW shearing is consistent with the orientation of steepplanar magnesite veins in between the shear zones (mean valueN134�, 62ºSW), interpreted as tension gashes (Quesnel et al., 2013),and with the orientation of folds developed at the expense ofmagnesite veins within the shear zones (Fig. 12e). Top-to-SWshearing is also consistent with the apparent top-to-west vs. top-to-south sense of shear observed along the ~WeE-trending VA1vs. ~NeS-trending VA2 section, respectively. The orientation of theellipsoid delineated by the large ovoid clast in the upper left cornerof Fig. 3a also fits with this interpretation. Along the VA1 section,the major shear zones with an apparent top-to-west sense of shearhave a mean apparent dip of 24.5� (Fig. 2c). Assuming that themean strike of these shear zones is N141� (i.e. the same as for themeter-scale shear planes) and taking into account the N082� strikeof the VA1 section, the mean true dip of the major shear zones canbe estimated at ~28�. Taking into account the likely ~6e7� west-ward slope of the basal contact of the nappe beneath the Vavoutopeninsula (Fig. 2b), the obliquity of the main shear zones withrespect to the boundaries of the serpentine sole is probably around22�. This obliquity is consistent with an interpretation of the shearzones as large C’-type shear bands (Berth�e et al., 1979; Passchierand Trouw, 1996) developed during strong tangential shear acrossthe serpentine sole. Smaller equivalent structures are visible inFigs. 12c, 13b and 13c. Finally, it should be noted that flat-lyingantithetic shear planes also occur in the eastern part of the VA1section (Figs. 2c and 12b). In Fig. 12b, they crosscut the mainsouthwest-dipping shear fabric, however their relationship withone of the major shear zones suggests that they may have devel-oped contemporaneously.

5. Discussion

5.1. The distribution of deformation in the Koniambo Massif

Fig. 14 summarizes the results of our structural analysis whileFig. 15 illustrates the possible relationships between the threestructural levels identified in the Koniambo Massif.

The Upper structural level is characterized by a very densenetwork of fractures filled with one or several types of serpentine.Antigorite and polygonal serpentine commonly form slickenfibersalong fault planes, which we used for analyzing fault slip data. Thebulk deformation associated with antigorite-bearing faults ischaracterized by a highly constrictional ellipsoid with l1 lyinghorizontally along a N102� trend. The bulk deformation associatedwith polygonal serpentine-bearing faults is characterized by anellipsoid in the flattening field with l3 lying subhorizontally along aN144� trend.

The Lower structural level coincides with the serpentine sole. As

Fig. 11. Field views of the reverse-slip fault zone with top-to-ENE kinematics shown inFig. 2a,b.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 61

Page 45: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

a rule in New Caledonia (e.g., Avias, 1967; Legu�er�e, 1976; Cluzelet al., 2012), the serpentine sole of the Koniambo Massif isintensely deformed. Breccias dominate while foliated serpentinites,interpreted as mylonitic rocks, form a distinct basal layer at least20 m thick. The sole has experienced pervasive non-coaxialdeformation with a top-to-SW sense of shear. Polygonal serpen-tine and magnesite are syn-kinematic phases with respect to thisdeformation. Along the VA1 section, major low-dipping shearzones are distributed with a ~100 m lengthscale (Fig. 2c). Asdescribed in Section 4.3.3, their mean obliquity with respect to thebasal contact of the nappe is estimated at ~22�. We interpret these

shear zones as C’-type shear bands developedwithin a thick zone ofstrong tangential shear that coincides with the serpentine sole(Figs. 14 and 15; see also Lahond�ere et al., 2012; their Figure 115). Inline with this interpretation, no equivalent shear zone is observedabove the serpentine sole.

The Intermediate structural level is characterized by the pres-ence of several meters-thick reverse-slip shear zones. Top-to-SWshear zones are synthetic to pervasive shearing along the serpen-tine sole and, likewise, involve polygonal serpentine andmagnesiteas syn-kinematic mineral phases. Therefore, although we had noopportunity to observe this contact in the field, we suppose that the

Fig. 12. Field views of the serpentine sole in the Vavouto peninsula, along the (a & b) VA1 and (c & d) VA2 sections (location in Fig. 2). Plg, polygonal serpentine; Mgs, magnesite. e)Lower hemisphere, equal-area projection showing the orientation, along the VA1 section, of large shear planes (shown as orange squares representing poles of planes), of planarmagnesite veins (shown as great circles) in between major shear zones, and of fold axes (shown as black dots) of folded magnesite veins within the shear zones. (For interpretationof the references to color in this figure legend, the reader is referred to the web version of this article.)

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6762

Page 46: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

shear zones root along the roof of the serpentine sole, the wholesole acting as a d�ecollement (Fig. 14 and, in Fig. 15, geometricrelationship ‘r1’). In Fig. 15, the roof of the sole is schematicallyshown as a distinct fault; in reality, it is more likely to coincide with

a zone across which the intensity of brecciation progressively di-minishes. In addition, some shear zones may root at deeper level(Fig. 15, relationship ‘r2’), for instance along the relatively sharpcontact between the foliated serpentinites and the overlying

Fig. 13. Field views (a to c) and microphotograph (d) illustrating the relationships between magnesite and deformation in the serpentine sole.

Fig. 14. Left, block-diagram depicting the style of deformation in the three structural levels identified in the Koniambo Massif. For the Upper structural level, the drawing shows thedeformation associated with polygonal serpentine, rather than with antigorite, because polygonal serpentine is also the main mineral phase associated with major structures in theother structural levels. Roadsite ‘620 m’ is shown at the top of the Intermediate structural level to highlight its location at lower elevation in comparison to the main area of faultmeasurements (see Fig. 2a). Right, synthesis of the results of strain inversion as a function of the structural level (vertically) and as a function of the main minerals associated withdeformation, which enable the determination of a sequence of events (horizontally).

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 63

Page 47: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

serpentinite breccias (Fig. 3b) and/or along the basal contact of thenappe. This possibility is suggested by the presence of a relativelylarge top-to-SW reverse-slip shear zone at the northern end of theVA2 section (Fig. 3a) and by the characteristics of the fault zone ofFig. 11 (see Section 4.3.2 and below).

The NEeSW horizontal spacing between two consecutive top-to-SW shear zones, SZ1 and SZ2, is ~450 m. This is slightly lessthan the thickness of the Upper structural level in the KoniamboMassif (Fig. 2b) and much less than the thickness of the PeridotiteNappe in the Massif du Sud (�1.5 km). This narrow spacing sug-gests that at least some of the shear zones do not cross the nappeup to the surface but remain confined to the Intermediate structurallevel. In fact, although rock exposures are abundant in the open pitsof the top of the Koniambo Massif, we have not observed any suchshear zone in the Upper structural level. Based on this indirectevidence, it is here suggested that the shear zones may connect toanother flat-lying d�ecollement located along the roof of the Inter-mediate structural level (Fig. 15, relationship ‘r3’). Like for the roofof the sole, this d�ecollement might not coincide with a distinct faultbut with a zone across which the reverse-slip displacement of theshear zones could be absorbed through diffuse faulting. In addition,a few major shear zones may ramp up to the surface (Fig. 15,relationship ‘r4’). No such shear zone is observed in the KoniamboMassif but one likely exists in the northern part of the nearbyKop�eto-Boulinda Massif (see Fig. 1 for location). It is highlighted bya north-dipping serpentinite sheet about 200 m thick that ema-nates from the serpentine sole of this massif, at elevations around200 m, and climbs among the peridotites up to an elevation of atleast 500 m (Maurizot et al., 1985; their Figure 4; Maurizot, 2007).The precise geometry and the kinematics of this probable shearzone remain to be documented.

Finally, top-to-NE shear zones also exist in the KoniamboMassif,as subsidiary structures in the serpentine sole and as conjugateshears with respect to the top-to-SW shear zones in the Interme-diate structural level. Because it roots along or slightly beneath thebase of the serpentine sole and likely offsets its roof (Fig. 2b), thereverse fault zone of Fig. 11 is probably a relatively late feature withrespect to pervasive top-to-SW shearing along the sole. The shearzone labeled ‘r2’, in Fig. 15, would be an equivalent synthetic shearzone.

5.2. The temporal evolution of deformation

Our analysis documents major variations in the orientation of

the principal strain axes. At least part of these variations reflect atemporal evolution (Fig. 14). This is the case within the Upperstructural level, where fault slip data associated with antigorite andpolygonal serpentine were collected in the same restricted area butyielded two strikingly different strain ellipsoids. Microstructuralobservations indicate that polygonal serpentine postdates anti-gorite (see Sections 4.1 and 4.3.1), in agreement with the timesequence identified by Ulrich (2010) in the serpentine sole of theKoniambo Massif. Hence, the Upper structural level keeps the re-cord of a temporal change from WNW-ESE horizontal stretching,during antigorite crystallization, to NWeSE horizontal shortening,during polygonal serpentine crystallization.

The origin of the other variations in orientation of the principalstrain axes is more difficult to assess because they involve the sameserpentine polymorph (polygonal serpentine) and occur acrossdistinct levels of the nappe. Therefore, the change from NWeSEshortening in the Upper structural level to NEeSW shearing in theserpentine sole could reflect a spatial rather than a temporal evo-lution, i.e. vertical strain partitioning at a specific evolutionary stageof the nappe. The ~ NeS direction of shortening recorded at anintermediate structural height, on roadsite ‘620 m’, may supportthis view. Nevertheless, the observed changes in shortening di-rection aremore likely to reflect a temporal evolution. A first reasonis that it seems difficult to conceive a tectonic setting in which thenappe would undergo horizontal shortening at right angle to itsdirection of displacement (assuming the latter is given by the di-rection of shear along the basal d�ecollement). A second reason isthe fact that, in the Intermediate structural level, the main shearzones, which accommodate NEeSW shortening, crosscut rockswith the same dense network of fractures as in the Upper structurallevel (Fig. 10a). This relation suggests that the event having pro-duced NEeSW shortening in the Intermediate structural level (andtop-to-SW shearing in the sole) is younger than the fracture sets ofthe Upper structural level.

Thus, we suspect that both the syn-antigorite and the syn-polygonal serpentine fault sets identified in the Upper structurallevel are older than the top-to-SW shear deformation observed inthe serpentine sole. This contrasts with the opinion of Legu�er�e(1976) who supposed that the various sets of fractures he identi-fied in the main mass of the nappe are all younger than thedeformation recorded by the sole. According to Legu�er�e (1976),these fracture sets provide evidence for a three-step temporalevolution involving an episode of ENE-WSW compression, then anepisode of NNW-SSE compression, then an episode of NNW-SSE

Fig. 15. Schematic cross-section depicting the vertical distribution of deformation within the Peridotite Nappe by the time top-to-SW shearing occurred along the serpentine sole.‘r1’ to ‘r4’ are geometric relationships discussed in the text. The thickness and the topography of the nappe are poorly constrained and are likely to have changed duringdeformation.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6764

Page 48: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

extension. This scenario applies to the Kop�eto-Boulinda andKoniambo Massifs, but the same events and the same chronology,with some variations in the orientation of the strain axes, are givenby Legu�er�e (1976) for the other klippes of the Peridotite Nappe heexamined, spread over the Grande Terre. The episode of NNW-SSEcompression, of fairly constant orientation across the Grande Terre(Legu�er�e, 1976), correlates well with the syn-polygonal serpentineevent we identified in the Upper structural level of the KoniamboMassif. The episode of NNW-SSE extension, which is more variablein orientation on the scale of the Grande Terre (extension is locallyWNW-ESE to WeE, cf. Legu�er�e, 1976), may correlate with the syn-antigorite event we identified. Finally, the episode of ENE-WSWcompression, locally modified to a NEeSW compression (Legu�er�e,1976), correlates with the syn-polygonal serpentine ± magnesiteevent we identified in the Intermediate structural level and theserpentine sole. If so, however, the chronology of events proposedby Legu�er�e (1976) differs strikingly from ours: his first event is ourlast one, and his last event is our first one.We ignore the reasons forthis disagreement, but note that Legu�er�e (1976) did not providedetails on the arguments he used for establishing his chronology.Our chronology is essentially based on the identification of asequence of minerals associated with deformation, from antigoriteto polygonal serpentine ± magnesite. This sequence is consistentwith a progressive lowering in temperature conditions (Ulrich,2010), which may reflect a modification of the geothermalgradient and/or the progressive reduction in thickness of the nappe(Fig. 15). As a possible explanation for our disagreement withLegu�er�e (1976), we might have failed to identify one or several lateepisodes of normal faulting that are reported from the sedimentaryseries surrounding the klippes of the Peridotite Nappe (Lagabrielleet al., 2005; Chardon and Chevillotte, 2006).

5.3. Relationships between deformation and serpentinization

The three structural levels identified in the Koniambo Massifshow strikingly different styles of deformation. These levelscorrelate fairly well with the subhorizontal lithological layering ofthe massif defined by Maurizot et al. (2002), which refers to vari-ations in the intensity of serpentinization. Hence, it is tempting toconsider that the intensity of serpentinization played a major rolein the way deformation has been distributed across the PeridotiteNappe. As implicitly considered in several studies (e.g., Avias, 1967;Cluzel et al., 2001), the massive serpentinites of the sole mightrepresent a layer of soft rocks having led to the individualization ofa major d�ecollement at the base of the nappe.

The Upper structural level displays a dense network ofserpentine-bearing fractures, many of which have been activated asfaults. Nevertheless, the bulk deformation associated with thesefaults is probably limited, as suggested by the fairly constantorientation of dunitic interlayers all across the Upper structurallevel (Maurizot et al., 2002). Moreover, as discussed in Section 5.2,the deformation recorded by the serpentine sole and the shearzones of the Intermediate structural level is likely younger thanmost fault sets of the Upper structural level. This supports thepicture of an almost rigid Upper structural level by the timeshearing was accumulating along the sole (cf. the strain-depth di-agram in Fig. 15). Could this behavior relate to a least degree of bulkserpentinization in the Upper structural level?

Unlike in some parts of the Massif du Sud (e.g., Orloff, 1968),pristine peridotites are virtually absent from the Koniambo Massif.The least serpentinized rocks are found in the highest ~200m of themassif. In the field, they are identified as they seem to retain agranoblastic texture, yet their dark green color is symptomatic of asignificant serpentine content. Under microscope, these samplesshow ‘mesh’ textures resulting from a partial replacement of olivine

by lizardite (Fig. 6b). Serpentine mesh textures are essentiallypseudomorphic, therefore peridotites with these textures usuallypreserve their pre-serpentinization fabrics (e.g., Wicks andWhittaker, 1977; Roum�ejon and Cannat, 2014), which also appliesin that case (Fig. 6b; for other examples from the Koniambo Massif,see Lahond�ere et al., 2012, their Figures 301 to 304 and 309 to 312).This suggests that the Upper structural level behaved almost rigidlyever since the early development of lizardite.

The mesh texture in Fig. 6b is classical (see also the examples ofLahond�ere et al., 2012) in the sense that lizardite forms connected‘mesh rims’ surrounding isolated olivine ‘mesh cores’ (e.g.,Roum�ejon and Cannat, 2014). This is a key observation becausetheoretical considerations (e.g., Handy et al., 1999) as well asexperimental results (Escartín et al., 2001) indicate that, with sucha texture, the peridotite should be nearly as weak as pure serpen-tinite. The rock in Fig. 6b contains ~25% serpentine (estimated fromimage analysis) whereas experimental evidence suggests that a~10e15% serpentine content is enough for achieving the largestpart of the strength drop with respect to the strength of a pureperidotite (Escartín et al., 2001). Therefore, it seems that the highlyuneven vertical distribution of syn to post-serpentinization defor-mation across the height of the Koniambo Massif does not relate toa parallel gradient in rock strength: within the range of observedvariations in the degree of serpentinization, all the rocks shouldhave approximately the same strength (cf. Escartín et al., 2001).

5.4. What promoted strain localization at lower levels of the nappe ?

The above discussion has shown that the degree of serpentini-zation is probably not the factor which led deformation to localizealong the sole of the nappe. A simple alternative may be consideredwhere the gradient in rock strength is not located within thePeridotite Nappe but arises from its juxtaposition against rocks ofthe substratum. In the Koniambo Massif, as for the other klippes ofthe northwestern coast, this substratum is essentially composed ofbasalts and dolerites (e.g., Guillon, 1975; Paris, 1981; Cluzel et al.,2001) (Fig. 3a). Such rocks are clearly stronger than serpentinites(e.g., Ildefonse et al., 2007) and, following Escartín et al. (2001), alsostronger than slightly serpentinized peridotites. Hence, a strengthprofile may be conceived where the whole Peridotite Nappe isweak, overlying a stronger substratum. Deformation is then ex-pected to concentrate in the part of the weak domain that is theclosest to the strong one, i.e. on the soft side of the main rheologicalboundary. This could account for strain localization along the soleof the nappe.

Where basalts and dolerites underlie the Peridotite Nappe, thethickness of the serpentine sole is great (~100e200 m). This is thecase in the Koniambo Massif and the other klippes of the north-western coast (e.g., Guillon, 1975). In contrast, smaller klippeslocated further northeast, around the axis of the island, have amuch thinner serpentine sole (~10e20 m) and a substratum madeof pervasively schistose fine-grained sediments with a low grademetamorphic overprint. Typical examples are the Tchingou Massif(e.g., Maurizot et al., 1985) and a series of kilometer-sized klippesaround the Ougne summit (Maurizot et al., 1989) (Fig. 1). Assumingthat the thickness of the sole depends essentially on the intensity ofshearing, and assuming that the total amount of shear, linked to thedisplacement of the nappe, is approximately constant over the areaunder consideration, this large difference in sole thickness isconsistent with the hypothesis that a lower vs. higher amount ofstrain is accommodated by the nappe when its substratum is madeof soft metasediments vs. strong mafic rocks, respectively. Thissupports our proposal that strain localization along the base of thenappe results from its juxtaposition against a stronger substratum.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 65

Page 49: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

6. Conclusions

Three structural levels have been identified in the KoniamboMassif. The Upper structural level, characterized by a densenetwork of fractures, keeps the record of at least two deformationevents, the first associated with antigorite (WNW-ESE extension),the second with polygonal serpentine (NWeSE compression). TheLower structural level is represented by the serpentine sole. Itconsists of massive tectonic breccias overlying a layer of myloniticserpentinites. The sole, which records pervasive tangential shearwith top-to-SW kinematics, can be interpreted as a d�ecollement atthe base of the nappe. The Intermediate structural level exposesseveral meters-thick conjugate shear zones accommodatingNEeSW shortening. Like the sole, these shear zones involvepolygonal serpentine and magnesite as the main syn-kinematicmineral phases. The shear zones likely root into the basald�ecollement, either along its roof or, occasionally, around its base.

Even the least altered peridotites, in the Upper structural level,contain so much serpentine that, according to theoretical andexperimental work, they should be nearly as weak as pure ser-pentinite. Hence, no strong vertical gradient in strength due tovariations in the degree of serpentinization is expected within theexposed part of the nappe. Strain localization along the serpentinesole probably results from the juxtaposition of the nappe, made ofweak serpentinized peridotites, against strongermafic rocks. In linewith this view, the serpentine sole is about ten times thinner whenthe substratum of the nappe is made of weak metasediments,compared to what it is when the substratum is made of basalts anddolerites.

Acknowledgments

GeoRessources contribution has been supported by ANR-10-LABX-21eLABEX RESSOURCES 21. This work benefited from fruit-ful discussions with Marc Ulrich, Dominique Cluzel and PierreMaurizot. We gratefully thank Sarah J. Titus and an anonymousreviewer for their constructive comments, and Toru Takeshita foreditorial handling.

References

Allmendinger, R.W., Cardozo, N., Fisher, D.M., 2012. Structural Geology Algorithms.Cambridge University Press, Cambridge, p. 289.

Angelier, J., 1994. Fault slip analysis and paleostress reconstruction. In: Hancock, P.L.(Ed.), Continental Deformation. Pergamon Press, Oxford, pp. 53e100.

Angelier, J., Mechler, P., 1977. Sur une m�ethode graphique de recherche des con-traintes principales �egalement utilisable en tectonique et en sismologie: lam�ethode des di�edres droits. Bull. Soc. G�eol. Fr. 19, 1309e1318.

Auzende, A.L., Daniel, I., Reynard, B., Lemaire, C., Guyot, F., 2004. High-pressurebehaviour of serpentine minerals: a Raman spectroscopic study. Phys. Chem.Min. 31, 269e277.

Avias, J., 1967. Overthrust structure of the main ultrabasic New Caledonian mas-sives. Tectonophysics 4, 531e541.

Bailey, W.R., Holdsworth, R.E., Swarbrick, R.E., 2000. Kinematic history of a reac-tivated oceanic suture: the mamonia complex suture zone, SW Cyprus. J. Geol.Soc. Lond. 157, 1107e1126.

Berth�e, D., Choukroune, P., Gapais, D., 1979. Orientations pr�ef�erentielles du quartz etorthogneissification progressive en r�egime cisaillant: l’exemple du cisaillementsud-armoricain. Bull. Min�er. 102, 265e272.

Boudier, F., Ceuleneer, G., Nicolas, A., 1988. Shear zones, thrusts and related mag-matism in the Oman ophiolite: initiation of thrusting on an oceanic ridge.Tectonophysics 151, 275e296.

Butt, C.R.M., Cluzel, D., 2013. Nickel laterite ore deposits: weathered serpentinites.Elements 9, 123e128.

Carrou�e, J.P., 1972. Carte g�eologique de la Nouvelle-Cal�edonie �a l’�echelle 1/50000,feuille Pouembout, p. 38. Bureau de Recherches G�eologiques et Mini�eres, mapsheet and explanatory notes.

C�el�erier, B., Etchecopar, A., Bergerat, F., Vergely, P., Arthaud, F., Laurent, P., 2012.Inferring stress from faulting: from early concepts to inverse methods. Tecto-nophysics 581, 206e219.

Chardon, D., Chevillotte, V., 2006. Morphotectonic evolution of the New Caledoniaridge (Pacific Southwest) from post-obduction tectonosedimentary record.

Tectonophysics 420, 473e491.Chevillotte, V., Chardon, D., Beauvais, A., Maurizot, P., Colin, F., 2006. Long-term

tropical morphogenesis of New Caledonia (Southwest Pacific): importance ofpositive epeirogeny and climate change. Geomorphology 81, 361e375.

Cluzel, D., Aitchison, J.C., Picard, C., 2001. Tectonic accretion and underplating ofmafic terranes in the late eocene intraoceanic fore-arc of New Caledonia(Southwest Pacific): geodynamic implications. Tectonophysics 340, 23e59.

Cluzel, D., Maurizot, P., Collot, J., Sevin, B., 2012. An outline of the geology of NewCaledonia; from Permian-Mesozoic Southeast Gondwanaland active margin toCenozoic obduction and supergene evolution. Episodes 35, 72e86.

Cluzel, D., Vigier, B., 2008. Syntectonic mobility of supergene nickel ores of NewCaledonia (Southwest Pacific). Evidence from faulted regolith and garnieriteveins. Resour. Geol. 58, 161e170.

Coleman, R., 1971. Plate tectonic emplacement of upper mantle peridotites alongcontinental edges. J. Geophys. Res. 76, 1212e1222.

Cook, C.A., Holdsworth, R.E., Styles, M.T., Pearce, J.A., 2000. Pre-emplacementstructural history recorded by mantle peridotites: an example from the lizardcomplex, SW England. J. Geol. Soc. Lond. 157, 1049e1064.

Coulton, A.J., Harper, G.D., O'Hanley, D.S., 1995. Oceanic versus emplacement ageserpentinization in the Josephine ophiolite: implications for the nature of theMoho at intermediate and slow spreading ridges. J. Geophys. Res. 100,22245e22260.

Dewey, J.F., Bird, J.M., 1971. Origin and emplacement of the ophiolite suite: appa-lachian ophiolites in Newfoundland. J. Geophys. Res. 76, 3179e3206.

Durney, D.W., Ramsay, J.G., 1973. Incremental strains measured by syntectoniccrystal growths. In: de Jong, K.A., Scholten, R. (Eds.), Gravity and Tectonics.Wiley, New York, pp. 67e96.

Escartín, J., Hirth, G., Evans, B., 2001. Strength of slightly serpentinized peridotites:implications for the tectonics of oceanic lithosphere. Geology 29, 1023e1026.

Federico, L., Crispini, L., Vigo, A., Capponi, G., 2014. Unravelling polyphase brittletectonics through multi-software fault-slip analysis: the case of the Voltri Unit,Western Alps (Italy). J. Struct. Geol. 68, 175e193.

Gapais, D., Cobbold, P.R., Bourgeois, O., Rouby, D., de Urreiztieta, M., 2000. Tectonicsignificance of fault-slip data. J. Struct. Geol. 22, 881e888.

Guillon, J.H., 1975. Les massifs p�eridotitiques de Nouvelle-Cal�edonie. M�emoires del’Office de la Recherche Scientifique et Technique Outre-Mer, Paris, vol. 76,p. 120.

Handy, M.R., Wissing, S.B., Streit, L.E., 1999. Frictional-viscous flow in mylonite withvaried bimineralic composition and its effect on lithospheric strength. Tecto-nophysics 303, 175e191.

Hirth, G., Guillot, S., 2013. Rheology and tectonic significance of serpentinite. Ele-ments 9, 107e113.

Ildefonse, B., Blackman, D.K., John, B.E., Ohara, Y., Miller, D.J., MacLeod, C.J., 2007.Oceanic core complexes and crustal accretion at slow-spreading ridges. Geology35, 623e626.

Lagabrielle, Y., Chauvet, A., Ulrich, M., Guillot, S., 2013. Passive obduction andgravity-driven emplacement of large ophiolitic sheets: the New Caledoniaophiolite (SW Pacific) as a case study? Bull. Soc. G�eol. Fr. 184, 545e556.

Lagabrielle, Y., Maurizot, P., Lafoy, Y., Cabioch, G., Pelletier, B., R�egnier, M., Wabete, I.,Calmant, S., 2005. Post-Eocene extensional tectonics in Southern New Caledo-nia (SW Pacific): insights from onshore fault analysis and offshore seismic data.Tectonophysics 403, 1e28.

Lahond�ere, D., Lesimple, S., Cagnard, F., Lahfid, A., Wille, G., Maurizot, P., 2012.Serpentinisation et fibrogen�ese dans les massifs de p�eridotite de Nouvelle-Cal�edonie. Bureau de Recherches G�eologiques et Mini�eres, p. 458. Public ReportBRGM/RP-60192-FR.

Lahond�ere, D., Maurizot, P., 2009. Typologie et protocole d'�echantillonnage desoccurrences naturelles d'amiante en Nouvelle-Cal�edonie. Bureau de RecherchesG�eologiques et Mini�eres, p. 164. Public Report BRGM/RP-57334-FR.

La�o-D�avila, D.A., Anderson, T.H., 2009. Kinematic analysis of serpentinite structuresand the manifestation of transpression in southwestern Puerto Rico. J. Struct.Geol. 31, 1472e1489.

Latham, M., 1977. On the geomorphology of northern and western New Caledonianultramafic massifs. In: International Symposium on Geodynamics in South-west Pacific, 27 August-2 September 1976, Noumea. Editions Technip, Paris,New Caledonia, pp. 235e244.

Legu�er�e, J., 1976. Des corr�elations entre la tectonique cassante et l’alt�erationsuperg�ene des p�eridotites de Nouvelle-Cal�edonie. Ph.D. thesis. Universit�e deMontpellier, France, p. 94.

Marrett, R., Allmendinger, R.W., 1990. Kinematic analysis of fault-slip data. J. Struct.Geol. 12, 973e986.

Maurizot, P., 2007. Cartographie d’aide �a l’am�enagement dans la zone Voh-Kon�e-Pouembout-Poya. Phase 1: Mise �a jour de la carte g�eologique sur les communesde Kon�e et Pouembout et bilan des donn�ees existantes. Bureau de RecherchesG�eologiques et Mini�eres, p. 76. Public Report BRGM/RP-54898-FR.

Maurizot, P., Eberl�e, J.M., Habault, C., Tessarolo, C., 1989. Carte g�eologique de laNouvelle-Cal�edonie �a l’�echelle 1/50000, feuille Pam-Ou�egoa, 2 �edition, p. 81.Bureau de Recherches G�eologiques et Mini�eres, map sheet and explanatorynotes.

Maurizot, P., Feigner, D., Paris, J.P., 1985. Donn�ees nouvelles sur les « fils de ser-pentinite » de Nouvelle-Cal�edonie. G�eol. Fr. 1, 61e67.

Maurizot, P., Lafoy, Y., Poup�ee, M., 2002. Cartographie des formations superficielleset des al�eas mouvements de terrain en Nouvelle-Cal�edonie, zone du Koniambo.Bureau de Recherches G�eologiques et Mini�eres, p. 45. Public Report BRGM/RP51624-FR.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e6766

Page 50: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Maurizot, P., Vend�e-Leclerc, M., 2009. Carte g�eologique de la Nouvelle-Cal�edonie �al’�echelle 1/500000. Direction de l’Industrie, des Mines et de l’Energie e Servicede G�eologie de Nouvelle-Cal�edonie, Bureau de Recherches G�eologiques etMini�eres.

Orloff, O., 1968. Etude g�eologique et g�eomorphologique des massifs d’ultrabasitescompris entre Houailou et Canala (Nouvelle-Cal�edonie). Ph.D. thesis. Universit�ede Montpellier, France, p. 189.

Paris, J.P., 1981. G�eologie de la Nouvelle-Cal�edonie, un essai de synth�ese. M�emoiresdu Bureau de Recherches G�eologiques Mini�eres 113, 274.

Passchier, C.W., Trouw, R.A.J., 1996. Microtectonics. Springer, Berlin, p. 289.Prinzhofer, A., Nicolas, A., Cassard, D., Moutte, J., Leblanc, M., Paris, J.P.,

Rabinovitch, M., 1980. Structures in the New Caledonia peridotites-gabbros:implications for oceanic mantle and crust. Tectonophysics 69, 85e112.

Quesnel, B., Gautier, P., Boulvais, P., Cathelineau, M., Maurizot, P., Cluzel, D.,Ulrich, M., Guillot, S., Lesimple, S., Couteau, C., 2013. Syn-tectonic, meteoricwater-derived carbonation of the New Caledonia peridotite nappe. Geology 41,1063e1066.

Roum�ejon, S., Cannat, M., 2014. Serpentinization of mantle-derived peridotites atmid-ocean ridges: mesh texture development in the context of tectonic exhu-mation. Geochem. Geophys. Geosys. 15, 2354e2379.

Schemmann, K., Unruh, J.R., Moores, E.M., 2007. Kinematics of Franciscan Complexexhumation: new insights from the geology of Mount Diablo, California. Geol.Soc. Am. Bull. 120, 543e555.

Schwartz, S., Guillot, S., Reynard, B., Lafay, R., Debret, B., Nicollet, C., Lanari, P.,Auzende, A.L., 2013. Pressureetemperature estimates of the lizardite/antigoritetransition in high pressure serpentinites. Lithos 178, 197e210.

Spandler, C., Rubatto, D., Hermann, J., 2005. Late Cretaceous-Tertiary tectonics of thesouthwest Pacific: insights from U-Pb sensitive, high-resolution ion microprobe(SHRIMP) dating of eclogite facies rocks from New Caledonia. Tectonics 24,TC3003.

Suhr, G., Cawood, P.A., 1993. Structural history of ophiolite obduction, Bay of Islands.Nfld. Geol. Soc. Am. Bull. 105, 399e410.

Tikoff, B., Wojtal, S.F., 1999. Displacement control of geologic structures. J. Struct.Geol. 21, 959e967.

Titus, S.J., Fossen, H., Pedersen, R.B., Vigneresse, J.L., Tikoff, B., 2002. Pull-apartformation and strike-slip partitioning in an obliquely divergent setting, LekaOphiolite, Norway. Tectonophysics 354, 101e119.

Titus, S.J., Maes, S.M., Benford, B., Ferr�e, E.C., Tikoff, B., 2011. Fabric development inthe mantle section of a paleo-transform fault and its effect on ophioliteobduction, New Caledonia. Lithosphere 3, 221e244.

Trescases, J.J., 1975. L’�evolution g�eochimique superg�ene des roches ultrabasiques enzone tropicale, formation des gisements nick�elif�eres de Nouvelle-Cal�edonie.M�emoires de l’Office de la Recherche Scientifique et Technique Outre-Mer,Paris, vol. 78, p. 259.

Twiss, R.J., Moores, E.M., 1992. Structural Geology. W.H. Freeman and Company,New York, p. 532.

Twiss, R.J., Unruh, J.R., 1998. Analysis of fault slip inversions: do they constrainstress or strain rate? J. Geophys. Res. 103, 12205e12222.

Ulrich, M., 2010. P�eridotites et serpentinites du complexe ophiolitique de la Nou-velle-Cal�edonie. Ph.D thesis. Universit�e de Nouvelle-Cal�edonie et Universit�eJoseph Fourier (Grenoble), France, p. 246.

Ulrich, M., Mu~noz, M., Guillot, S., Cathelineau, M., Picard, C., Quesnel, B., Boulvais, P.,Couteau, C., 2014. Dissolutioneprecipitation processes governing the carbon-ation and silicification of the serpentinite sole of the New Caledonia ophiolite.Contrib. Mineral. Pet. 167, 952.

Ulrich, M., Picard, C., Guillot, S., Chauvel, C., Cluzel, D., Meffre, S., 2010. Multiplemelting stages and refertilization as indicators for ridge to subduction forma-tion: the New Caledonia ophiolite. Lithos 115, 223e236.

Wicks, F.J., Whittaker, E.J.W., 1977. Serpentine textures and serpentinization. Can.Mineral. 15, 459e488.

B. Quesnel et al. / Journal of Structural Geology 85 (2016) 51e67 67

Page 51: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Conditions of Magnesite formation

in ophiolites

Quesnel, B., Gautier, P., Boulvais, P., Cathelineau, M.,Maurizot, P.., Cluzel, D., Ulrich, M., Guillot, S.,

Lesimple, S., Couteau C. (2013)

Syn-tectonic, meteoric water-derived carbonation of the New Caledonia peridotite nappe,

Geology 41, 10 (2013) 1063-1066

- 3 -

Page 52: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

GEOLOGY | October 2013 | www.gsapubs.org 1063

INTRODUCTIONCarbonation of ultramafi c rocks is the process

by which CO2-bearing fl uids react with olivine and/or serpentine to form magnesite (MgCO3) (e.g., Klein and Garrido, 2011). Based on stable isotope and structural evidence, Kelemen et al. (2011) recently showed that present-day carbon-ation of the Oman ophiolite is due to downward infi ltration of meteoric waters in the absence of signifi cant tectonic activity. Other stable isotope studies have also established the meteoric origin of the fl uids from which magnesite has formed in a number of ophiolite occurrences (Barnes et al., 1973; Jedrysek and Halas, 1990; Fallick et al., 1991; Gartzos, 2004; Jurkovic et al., 2012; Oskierski et al., 2013). Some of these ophiolites include laterites and associated iron-nickel ore

deposits capping the ultramafi c rocks (e.g., El-iopoulos et al., 2012).

The main ophiolite of New Caledonia (south-west Pacifi c Ocean), referred to as the peridotite nappe, has also undergone intense laterization since its emergence. This has led to supergene nickel ore formation, a process which implies a well-drained percolation system through the peridotites (Trescases, 1975). Recently exposed outcrops in the Koniambo massif (Fig. 1A) show large surfaces of the serpentine sole that forms the base of the nappe (Figs. 1B and 2), providing unprecedented access to fresh samples. Numer-ous magnesite veins are observed along these outcrops, attesting to widespread carbonation.

Here we present oxygen and carbon isotope compositions of the magnesite veins and argue

that they originate from meteoric water. Further-more, in contrast with the situation depicted in Oman (Kelemen et al., 2011), many veins ap-pear to have formed syn-tectonically. This leads us to infer potential genetic links between lateri-zation, carbonation, and tectonics.

GEOLOGICAL SETTINGNew Caledonia lies 2000 km east of Australia.

About 40% of the island’s surface consists of pe-ridotite. Peridotites overlie rock units of the Nor-folk Ridge microcontinent with a sub-horizontal contact marked by a strongly deformed serpen-tine sole (Avias, 1967). This geometry results from the southwestward obduction of the perido-tite nappe, initially rooted in the Loyalty Basin, sometime between ca. 37 and 27 Ma (Cluzel et al., 2001, 2012; Paquette and Cluzel, 2007).

On top of the nappe, laterites have devel-oped at the expense of the peridotites (Tres-cases, 1975). Several planation surfaces attest to distinct episodes of weathering since before ca. 20 Ma (Latham, 1986; Chevillotte et al., 2006; Sevin et al., 2012). This is consistent with biogeographic and phylogenetic studies indicat-ing that New Caledonia was aerially exposed in the Late Oligocene (Grandcolas et al., 2008).

Magnesite is widespread in New Caledonia and occurs as veins within the serpentine sole of the peridotite nappe, and as nodular heaps in recent alluvial deposits and present-day soils. Since Glasser (1904), the origin of the veins is

*E-mails: [email protected]; [email protected]; [email protected]; [email protected].

Syn-tectonic, meteoric water–derived carbonation of the New Caledonia peridotite nappeBenoît Quesnel1*, Pierre Gautier1*, Philippe Boulvais1*, Michel Cathelineau2*, Pierre Maurizot3, Dominique Cluzel4, Marc Ulrich2,5, Stéphane Guillot5, Stéphane Lesimple6, and Clément Couteau7

1Géosciences Rennes, Université Rennes 1, UMR 6118 CNRS, 35042 Rennes Cedex, France2Georessources, Université de Lorraine, UMR CNRS 7359, CREGU, 54506 Vandoeuvre-lès-Nancy, France3Bureau de Recherches Géologiques et Minières, 98845 Noumea, New Caledonia4Pole Pluridisciplinaire de la Matière et de l’Environnement, Université de la Nouvelle-Calédonie, 98851 Nouméa, New Caledonia5ISTerre CNRS, Université Grenoble 1, 38041 Grenoble, France6Service Géologique de la Nouvelle-Calédonie, Direction de l’Industrie, des Mines et de l’Energie, 98851 Nouméa, New Caledonia7Service Géologique, Koniambo Nickel SAS, 98883 Voh, New Caledonia

ABSTRACTExceptional outcrops recently exposed in the Koniambo massif allow the study of the ser-

pentine sole of the peridotite nappe of New Caledonia (southwest Pacifi c Ocean). Many mag-nesite veins are observed, with characteristics indicating that they were emplaced during per-vasive top-to-the-southwest shear deformation. The oxygen isotope composition of magnesite is homogeneous (27.4‰ < δ18O < 29.7‰), while its carbon isotope composition varies widely (−16.7‰ < δ13C < −8.5‰). These new data document an origin of magnesite from meteoric fl uids. Laterization on top of the peridotite nappe and carbonation along the sole appear to represent complementary records of meteoric water infi ltration. Based on the syn-kinematic character of magnesite veins, we propose that syn-laterization tectonic activity has enhanced water infi ltration, favoring the exportation of leached elements like Mg, which has led to wide-spread carbonation along the serpentine sole. This calls for renewed examination of other magnesite-bearing ophiolites worldwide in order to establish whether active tectonics is com-monly a major agent for carbonation.

GEOLOGY, October 2013; v. 41; no. 10; p. 1063–1066; Data Repository item 2013295 | doi:10.1130/G34531.1 | Published online 30 July 2013

© 2013 Geological Society of America. For permission to copy, contact Copyright Permissions, GSA, or [email protected].

Figure 1. A: Simplifi ed geological map of New Caledonia, south-west Pacifi c Ocean. B: Geological map of southwestern margin of Koniambo massif, adapted from Maurizot et al. (2002). BMS, VAV, and CONV indicate magnesite sampling sites. Laterites shown on this map belong to a planation surface that is younger than those having led to thicker laterites at higher levels of Koniambo massif (at elevations between ~400 and ~800 m; see text) (Latham, 1986; Chevillotte et al., 2006; Chardon and Chevillotte, 2006).

Page 53: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

1064 www.gsapubs.org | October 2013 | GEOLOGY

supposed supergene, possibly linked to the lat-erization process.

OBSERVATIONS AND SAMPLINGThe Koniambo massif is one of the klippes

of the peridotite nappe located along the west coast (Fig. 1A). Recently, Koniambo Nickel SAS initiated a large industrial site for nickel production. As a result, new outcrops of excep-tional quality and size have been created in the serpentine sole of this massif (Figs. 1B and 2).

In the serpentine sole, rocks are strongly de-formed, either schistose or intensely brecciated. A dense network of meter-thick shallow-dipping shear zones attests to pervasive non-coaxial de-formation with a top-to-the-southwest sense of shear. Magnesite essentially occurs as veins, up to ~30 cm thick and irregularly distributed. Two main vein types are recognized (Figs. 3A–3C; Fig. DR1 in the GSA Data Repository1). Type 1 veins are located within or along the margins of the main shear zones. Open to tight drag fold-ing of some of these veins indicates that they formed during, or possibly before, shearing. Type 2 veins are steeper and occasionally cross-cut by the shallow-dipping shear zones, demon-strating that they do not represent younger struc-tures. The obliquity of these veins with respect to the shear zones (Fig. 3D) and the local occur-rence of magnesite as coarse fi bers orthogonal to vein walls (Fig. 3C) are consistent with their interpretation as tension gashes opened during top-to-the-southwest shearing.

Both vein types have been sampled along two cross sections (BMS and VAV) located in the Vavouto peninsula, just above the basal contact of the peridotite nappe (Fig. 1B; Table DR1 in the Data Repository). Samples were also col-lected in other highly serpentinized zones in the Koniambo massif (sample CONV, Fig. 1B) and in the Kopeto massif ~50 km to the southeast (samples NEP and GAIACS).

CARBON AND OXYGEN ISOTOPE DATAIsotopic analyses were performed at the

stable isotope laboratory of the University of Rennes 1, France. Samples were fi nely crushed in a boron carbide mortar and reacted with an-hydrous phosphoric acid at 75 °C for 24 h. The liberated CO2 was analyzed on a VG SIRA 10 triple collector mass spectrometer. The experi-mental fractionation coeffi cient between mag-nesite and CO2 is αCO2-magnesite = 1.009976 at 75 °C (Das Sharma et al., 2002). In the absence of a magnesite standard, in-lab calcite standard samples were analyzed together with the mag-nesite samples under identical conditions in or-der to control the general reliability of the pro-tocol. The analytical uncertainty is estimated at 0.3‰ for oxygen and 0.2‰ for carbon.

1GSA Data Repository item 2013295, oxygen and carbon isotope data, and additional fi eld observations, is available online at www.geosociety.org/pubs/ft2013.htm, or on request from [email protected] or Documents Secretary, GSA, P.O. Box 9140, Boulder, CO 80301, USA.

Figure 2. Field view along part of BMS cross section (located in Fig. 1B), illustrating ex-ceptional size and freshness of outcrops recently opened in serpentine sole of Koniambo massif.

Figure 3. Field observations within serpentine sole of Koniambo massif. A–C: Field views illustrating relations between deformation and magnesite veins. Numbers 1 and 2 refer to the two main vein types as described in text. Two ellipses in B show sites where folding of a “type 1” vein is well visible. In C, magnesite occurs as coarse fi bers suborthogonal to walls of this “type 2” vein (sample BMS Gio 9). D: Stereogram showing orientation, along BMS cross section, of major shear zones (shown as yellow squares representing poles of planes), of fold axes in the case of type 1 folded veins (shown as black dots), and of magnesite veins occurring in between shear zones (type 2 veins, shown as great circles).

Page 54: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

GEOLOGY | October 2013 | www.gsapubs.org 1065

Results are presented in Figure 4 and Ta-ble DR1. All magnesite samples display com-parable and high oxygen isotope values, irre-spective of their structural position or sampling site (27.4‰ < δ18O < 29.7‰); the carbon iso-tope composition is highly variable and nega-tive (−16.7‰ < δ13C < −8.5‰). Focusing on the samples from the Vavouto peninsula (BMS and VAV), the δ13C and δ18O values do not show any correlation.

METEORIC ORIGIN OF CARBONATIONThe isotopic compositions of magnesite re-

fl ect the conditions at which fl uid/rock interac-tions occurred. As the δ18O values are homoge-neous, the physical conditions of carbonation, δ18O value and temperature, were approximately constant in the fl uid. Based on the two arguments to follow, we suggest that a strong interaction with meteoric fl uids is the most likely process that led to carbonation along the serpentine sole. Firstly, although not precisely constrained, the carbon isotope fractionation between magnesite and CO2 is known to be positive (Oskierski et al., 2013), therefore magnesite is expected to display higher δ13C values than the fl uid from which it precipitated. As a result, the largely negative δ13C values in the magnesite veins rule out a seawater origin of the fl uid because carbon dissolved in seawater has a δ13C value close to 0‰. Secondly, as shown in Figure 4, our data set compares well with data from the literature on magnesite veins hosted by ultramafi c rocks and for which a meteoric origin of the fl uids has been proposed. Our data show a more restricted range in δ18O values centered on the right side of the literature data cloud, the Oman data ex-cluded. This observation may refl ect a differ-ence in the initial δ18O value of rainwater due to distinct paleogeographic position or a slightly different temperature of formation. The higher δ18O values from Oman probably refl ect the fact that veins there were formed from spring wa-ters, under lower temperatures (Kelemen et al., 2011). Our data show a large spread of δ13C val-ues that suggests at least two sources of carbon. The highest δ13C values are close to those from Oman, which is consistent with an origin of CO2 from the atmosphere. The lowest δ13C values, around −15‰, point to an organic carbon contri-bution, either from surface soils or from a deep-seated source such as methane liberated from sediments buried below the peridotite nappe.

Under high fl uid/rock ratios such as those in-ferred here because of the large size of magne-site veins and their abundance, the δ18O value of magnesite is a function of temperature and of the δ18O value of infi ltrated water. To date, no isoto-pic data are available for meteoric precipitations in New Caledonia. We use a δ18O range between −1‰ and −7‰, which corresponds to values of rainwater on isolated islands at inter-tropical lati-tude and low elevation (AIEA database; http://

www-naweb.iaea.org/napc/ih/IHS_resources_isohis.html#wiser). Using the recent study of Chacko and Deines (2008) on oxygen fraction-ation between magnesite and water, a tempera-ture range of 38–77 ºC can be estimated for the formation of the magnesite veins. This is consis-tent with heating of the meteoric waters while they were carried down to the serpentine sole.

LINKS BETWEEN LATERIZATION, CARBONATION, AND TECTONICS

The record of meteoric waters through car-bonation along the serpentine sole implies that water circulated downward through the perido-tite pile. An effi cient drainage system has likely been provided by the dense network of fractures that characterizes the New Caledonia perido-tites. This network is also recognized to have played a major role in peridotite weathering and the distribution of nickel ore (e.g., Leguéré, 1976). Hence, laterization on top of the perido-tite nappe and carbonation along the serpentine sole may correspond to complementary records of meteoric water infi ltration, as anticipated by Glasser (1904). In practice, laterization involves the leaching of large amounts of magnesium, a highly mobile ion that can be viewed as a tracer of fl uid circulation from the surface down to the serpentine sole where it precipitated to form magnesite veins. Correlatively, nickel, which is less mobile, has accumulated at the base of the lateritic profi le (e.g., Trescases, 1975). Through-out New Caledonia, the richest nickel ores are associated with a couple of planation surfaces associated with laterites up to 30 m thick (e.g., Chevillotte et al., 2006). They cap the Konia-mbo massif at elevations between ~400 and ~800 m, in agreement with the island-scale mean elevation of 640 m reported by Chevil-lotte et al. (2006). The outcrops of the Vavou-to peninsula, where most of our magnesite

samples come from, lie near sea level, and so does the basal contact of the peridotite nappe around much of the Koniambo massif. There-fore, downward infi ltration of meteoric waters likely occurred across a vertical distance of at least ~600 m before magnesite formed along the serpentine sole. Greater vertical distances are also possible because older laterite-bearing planation surfaces are locally preserved at eleva-tions up to ~1250 m in the nearby Kopeto massif and in southern New Caledonia (Latham, 1986; Chevillotte et al., 2006). This agrees with the above temperature range of ~38–77 ºC, which is consistent with fl uids in thermal equilibrium with host rocks at depths around 0.5–2.5 km, us-ing ~25 ºC as the initial temperature and 20–30 ºC/km for the geothermal gradient.

Because active slip typically increases the per-meability of faults, water drainage through the peridotites could have been enhanced during ac-tive faulting. Indirect evidence for deformation-assisted fl uid circulations across the peridotite pile is provided by the syn-kinematic character of the studied magnesite veins along the sole (Fig. 3; see also the Data Repository) and the ob-servation that at least some of the nickel miner-alizations underlying laterites developed during brittle deformation (Cluzel and Vigier, 2008; our own observations in the Koniambo massif).

The syn-kinematic magnesite veins of the Koniambo massif have been emplaced dur-ing top-to-the-southwest shearing deformation. Southwestward shearing recorded along the basal contact of the peridotite nappe may refl ect obduc-tion (e.g., Cluzel et al., 2001) or post-obduction reactivation of the contact as a southwest-dip-ping extensional detachment (Lagabrielle and Chauvet, 2008). Deformation occurred sometime between ca. 37 Ma and 27 Ma if related to ob-duction, or later, but before ca. 20 Ma, if related to post-obduction northeast-southwest extension

20 25 30−20

−18

−16

−14

−12

−10

−8

−6

BMS

NEP

VAV

CONV

GAIACS Balkans: Fallick et al., 1991

Greece: Gartzos, 2004

Poland: Jedrysek and Halas, 1990

δ18O vs. SMOW (‰)

Balkans: Jurkovic et al., 2012

−4

35

Oman: Kelemen et al., 2011

W California: Barnes et al., 1973

E Australia: Oskierski et al., 2013

δ13C

vs.

PDB

(‰)

Figure 4. δ13C versus δ18O diagram of magnesite veins hosted by ultra-mafi c rocks in different regions worldwide, for which a meteoric origin of fl uids has been pro-posed. Black and gray symbols are from this study and the literature, respectively. The results are given versus SMOW (standard mean ocean water) for δ18O and ver-sus PDB (Peedee belem-nite) for δ13C. BMS, VAV—Vavouto peninsula; CONV—Koniambo mas-sif; NEP and GAIACS—Kopeto massif.

Page 55: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

1066 www.gsapubs.org | October 2013 | GEOLOGY

(Chardon and Chevillotte, 2006). The main lat-erites of New Caledonia were also formed before ca. 20 Ma (Chevillotte et al., 2006; Sevin et al., 2012). Hence, available time constraints make it possible that carbonation and laterization oc-curred at the same time, during tectonic activity.

As a result, we propose that syn-laterization tectonic activity enhanced water infi ltration and played a major role in the exportation of leached elements like Mg, leading to widespread car-bonation along the serpentine sole.

POTENTIAL IMPLICATIONS FOR OTHER CARBONATED OPHIOLITES

Syn-tectonic carbonation along the serpentine sole of the New Caledonia ophiolite contrasts directly with the well-documented case of post-tectonic subsurface carbonation of the Oman ophiolite (Kelemen et al., 2011). Studies docu-menting meteoric water–derived magnesite in other ophiolite occurrences lack a structural de-scription that would allow the syn- versus post-tectonic character of carbonation to be evaluated (Barnes et al., 1973; Jedrysek and Halas, 1990; Fallick et al., 1991; Gartzos, 2004; Jurkovic et al., 2012; Oskierski et al., 2013). Nevertheless, syn-laterization tectonically driven carbonation of ultramafi c rocks, as proposed here for New Caledonia, may have occurred in other areas worldwide. For instance, the ophiolites of the Dinaric-Hellenic segment of the Alpine orogen include (1) large volumes of magnesite origi-nated from meteoric water (e.g., Gartzos, 2004; Jurkovic et al., 2012), (2) laterites capping the ultramafi c rocks, with iron-nickel ore deposits (e.g., Eliopoulos et al., 2012), (3) various time constraints showing that obduction occurred in the Late Jurassic, and (4) the unconformity of Late Jurassic sediments on at least some of the laterites (Robertson et al., 2012). These features strongly suggest that laterization occurred dur-ing obduction, which opens the possibility that carbonation occurred simultaneously, fostered by tectonic activity. This pleads for renewed examination of the Dinaric-Hellenic and other carbonated ophiolites worldwide in order to es-tablish whether active tectonics is commonly a major agent for carbonation.

ACKNOWLEDGMENTSThanks are due to Emmanuel Fritsch (Institut de

Recherche pour le Développement) for his help dur-ing preliminary fi eld work in the course of the Centre National de Recherche Technologique sur le Environ-ment program. Kerry Gallagher improved the English. Peter Kelemen and two anonymous reviewers helped us in clarifying some aspects of the paper.

REFERENCES CITEDAvias, J., 1967, Overthrust structure of the main ultra-

basic New Caledonian massives: Tectonophys-ics, v. 4, p. 531–541, doi:10.1016/0040-1951(67)90017-0.

Barnes, I., O’Neil, J.R., Rapp, J.B., and White, D.E., 1973, Silica-carbonate alteration of serpentine:

Wall rock alteration in mercury deposits of the California Coast Range: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 68, p. 388–398, doi:10.2113/gsecongeo.68.3.388.

Chacko, T., and Deines, P., 2008, Theoretical calcu-lation of oxygen isotope fractionation factors in carbonate systems: Geochimica et Cosmochi-mica Acta, v. 72, p. 3642–3660, doi:10.1016/j.gca.2008.06.001.

Chardon, D., and Chevillotte, V., 2006, Morpho tec-tonic evolution of the New Caledonia ridge (Pacifi c Southwest) from post-obduction tec-tonosedimentary record: Tectonophysics, v. 420, p. 473–491, doi:10.1016/j.tecto.2006.04.004.

Chevillotte, V., Chardon, D., Beauvais, A., Maurizot, P., and Colin, F., 2006, Long-term tropical morphogenesis of New Caledonia (Southwest Pacifi c): Importance of positive epeirogeny and climate change: Geomorphology, v. 81, p. 361–375, doi:10.1016/j.geomorph.2006.04.020.

Cluzel, D., and Vigier, B., 2008, Syntectonic mobil-ity of supergene nickel ores of New Caledonia (Southwest Pacifi c): Evidence from faulted regolith and garnierite veins: Resource Geol-ogy, v. 58, p. 161–170, doi:10.1111/j.1751-3928.2008.00053.x.

Cluzel, D., Aitchison, J.C., and Picard, C., 2001, Tec-tonic accretion and underplating of mafi c ter-ranes in the Late Eocene intraoceanic fore-arc of New Caledonia (Southwest Pacifi c): Geo-dynamic implications: Tectonophysics, v. 340, p. 23–59, doi:10.1016/S0040-1951(01)00148-2.

Cluzel, D., Jourdan, F., Meffre, S., Maurizot, P., and Lesimple, S., 2012, The metamorphic sole of New Caledonia ophiolite: 40Ar/39Ar, U-Pb, and geochemical evidence for subduction inception at a spreading ridge: Tectonics, v. 31, TC3016, doi:10.1029/2011TC003085.

Das Sharma, S., Patil, D.J., and Gopalan, K., 2002, Temperature dependence of oxygen isotope frac-tionation of CO2 from magnesite–phosphoric acid reaction: Geochimica et Cosmochimica Acta, v. 66, p. 589–593, doi:10.1016/S0016-7037(01)00833-X.

Eliopoulos, D.G., Economou-Eliopoulos, M., Apos-tolikas, A., and Golightly, J.P., 2012, Geochem-ical features of nickel-laterite deposits from the Balkan Peninsula and Gordes, Turkey: The genetic and environmental signifi cance of arse-nic: Ore Geology Reviews, v. 48, p. 413–427, doi:10.1016/j.oregeorev.2012.05.008.

Fallick, A.E., Ilich, M., and Russell, M.J., 1991, A sta-ble isotope study of the magnesite deposits asso-ciated with the Alpine-type ultramafi c rocks of Yugoslavia: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 86, p. 847–861, doi:10.2113/gsecongeo.86.4.847.

Gartzos, E., 2004, Comparative stable isotopes study of the magnesite deposits of Greece: Bulletin of the Geological Society of Greece, v. 36, p. 196–203.

Glasser, E., 1904, Rapport à M le Ministre des Colo-nies sur les richesses minérales de la Nouvelle-Calédonie: Paris, Annales des Mines, 560 p.

Grandcolas, P., Murienne, J., Robillard, T., Desutter-Grandcolas, L., Jourdan, H., Guilbert, E., and Deharveng, L., 2008, New Caledonia: A very old Darwinian island?: Philosophical Transactions of the Royal Society of London, Series B, v. 363, p. 3309–3317, doi:10.1098/rstb.2008.0122.

Jedrysek, M.O., and Halas, S., 1990, The origin of magnesite deposits from the Polish Foresudetic Block ophiolites: Preliminary δ13C and δ18O investigations: Terra Nova, v. 2, p. 154–159, doi:10.1111/j.1365-3121.1990.tb00057.x.

Jurkovic, I., Palinkas, L.A., Garasic, V., and Strmic Palinkas, S., 2012, Genesis of vein-stockwork cryptocrystalline magnesite from the Dinaride ophiolites: Ofi oliti, v. 37, p. 13–26.

Kelemen, P.B., Matter, J., Streit, E.E., Rudge, J.F., Curry, W.B., and Blusztajn, J., 2011, Rates and mechanisms of mineral carbonation in peridotite: Natural processes and recipes for enhanced, in situ CO2 capture and storage: Annual Review of Earth and Planetary Sciences, v. 39, p. 545–576, doi:10.1146/annurev-earth-092010-152509.

Klein, F., and Garrido, C.J., 2011, Thermodynamic constraints on mineral carbonation of serpen-tinized peridotite: Lithos, v. 126, p. 147–160, doi:10.1016/j.lithos.2011.07.020.

Lagabrielle, Y., and Chauvet, A., 2008, The role of extensional tectonics in shaping Cenozoic New-Caledonia: Bulletin de la Société Géologique de France, v. 179, p. 315–329, doi:10.2113/gssgfbull.179.3.315.

Latham, M., 1986, Altération et pédogenèse sur roches ultrabasiques en Nouvelle-Calédonie: Editions de l’ORSTOM (Offi ce de la Recherche Sci-entifi que et Technique Outre-Mer), Collection Etudes et Theses, 331 p.

Leguéré, J., 1976, Des corrélations entre la tectonique cassante et l’altération supergène des péridotites de Nouvelle Calédonie [Ph.D. thesis]: Montpel-lier, France, Université du Languedoc, 95 p.

Maurizot, P., Lafoy, Y., and Poupée, M., 2002, Car-tographie des formations superfi cielles et des aléas mouvements de terrain en Nouvelle-Calé-donie, Zone du Koniambo: Bureau de Recher-ches Géologiques et Minières Public Report RP51624-FR, 45 p.

Oskierski, H.C., Bailey, J.G., Kennedy, E.M., Jacob-sen, G., Ashley, P.M., and Dlugogoski, B.Z., 2013, Formation of weathering-derived magne-site deposits in the New England Orogen, New South Wales, Australia: Implications from min-eralogy, geochemistry and genesis of the Attunga magnesite deposit: Mineralium Deposita, v. 48, p. 525–541, doi:10.1007/s00126-012-0440-5.

Paquette, J.L., and Cluzel, D., 2007, U-Pb zircon dat-ing of post-obduction volcanic-arc granitoids and a granulite-facies xenolith from New Caledonia: Inference on Southwest Pacifi c geodynamic mod-els: International Journal of Earth Sciences, v. 96, p. 613–622, doi:10.1007/s00531-006-0127-1.

Robertson, A.H.F., Trivic, B., Deric, N., and Bucur, I.I., 2012, Tectonic development of the Vardar ocean and its margins: Evidence from the Re-public of Macedonia and Greek Macedonia: Tectonophysics, v. 595–596, p. 25–54, doi:10.1016/j.tecto.2012.07.022.

Sevin, B., Ricordel-Prognon, C., Quesnel, F., Cluzel, D., Lesimple, S., and Maurizot, P., 2012, First palaeomagnetic dating of ferricrete in New Caledonia: New insight on the morphogenesis and palaeoweathering of ‘Grande Terre’: Terra Nova, v. 24, p. 77–85, doi:10.1111/j.1365-3121.2011.01041.x.

Trescases, J.J., 1975, L’évolution géochimique super-gène des roches ultrabasiques en zone tropi-cale: Formation des gisements nickélifères de Nouvelle-Calédonie: Mémoires ORSTOM (Of-fi ce de la Recherche Scientifi que et Technique Outre-Mer), v. 78, 259 p.

Manuscript received 28 February 2013Revised manuscript received 20 May 2013Manuscript accepted 22 May 2013

Printed in USA

Page 56: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 57: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Conditions of Magnesite formation

in ophiolites

Ulrich M., Munoz, M. Guillot, S., Cathelineau, M., Picard, C., Quesnel, B.Boulvais , P., et Couteau C. (2014)

Dissolution–precipitation processes governing the carbonation and silicification of the serpentinite sole of the New

CaledoniaOphiolite

Contrib Mineral Petrol (2014) 167:952 DOI 10.1007/s00410-013-0952-8

- 4 -

Page 58: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

ORIGINAL PAPER

Dissolution–precipitation processes governing the carbonationand silicification of the serpentinite sole of the New Caledoniaophiolite

Marc Ulrich • Manuel Munoz • Stephane Guillot •

Michel Cathelineau • Christian Picard • Benoit Quesnel •

Philippe Boulvais • Clement Couteau

Received: 22 July 2013 / Accepted: 3 December 2013 / Published online: 18 January 2014

� Springer-Verlag Berlin Heidelberg 2014

Abstract The weathering of mantle peridotite tectoni-

cally exposed to the atmosphere leads commonly to natural

carbonation processes. Extensive cryptocrystalline mag-

nesite veins and stock-work are widespread in the ser-

pentinite sole of the New Caledonia ophiolite. Silica is

systematically associated with magnesite. It is commonly

admitted that Mg and Si are released during the laterization

of overlying peridotites. Thus, the occurrence of these

veins is generally attributed to a per descensum mechanism

that involves the infiltration of meteoric waters enriched in

dissolved atmospheric CO2. In this study, we investigate

serpentinite carbonation processes, and related silicifica-

tion, based on a detailed petrographic and crystal chemical

study of serpentinites. The relationships between serpen-

tine and alteration products are described using an original

method for the analysis of micro-X-ray fluorescence ima-

ges performed at the centimeter scale. Our investigations

highlight a carbonation mechanism, together with precipi-

tation of amorphous silica and sepiolite, based on a dis-

solution–precipitation process. In contrast with the per

descensum Mg/Si-enrichment model that is mainly con-

centrated in rock fractures, dissolution–precipitation pro-

cess is much more pervasive. Thus, although the texture of

rocks remains relatively preserved, this process extends

more widely into the rock and may represent a major part

of total carbonation of the ophiolite.

Keywords Serpentine � Magnesite � Carbonation �Silicification � New Caledonia ophiolite

Introduction

Carbon dioxide is currently one of the primary greenhouse

gases having an impact on global warming. Therefore,

numerous recent studies have focused on the potential of

sequestration of CO2 by mineral carbonation, either

through ex situ (e.g., Bobicki et al. 2012; Power et al. 2011;

Renforth et al. 2011; Balucan and Dlugogorski 2013;

Harrison et al. 2013) or in situ processes (e.g., Cipolli et al.

2004; Hansen et al. 2005; Teir et al. 2007, 2009; Andreani

Communicated by J. Hoefs.

Electronic supplementary material The online version of thisarticle (doi:10.1007/s00410-013-0952-8) contains supplementarymaterial, which is available to authorized users.

M. Ulrich (&) � M. Cathelineau

Laboratoire Georessources, CNRS, UMR 7566,

Universite de Lorraine, Nancy, France

e-mail: [email protected]; [email protected]

M. Ulrich � M. Munoz � S. Guillot

Institut des Sciences de la Terre, CNRS, UMR 5275,

Universite de Grenoble 1, Grenoble, France

Present Address:

M. Ulrich

IPGS-EOST, UMR 7516, Universite de Strasbourg,

Strasbourg, France

C. Picard

Laboratoire Chrono-environnement, UMR 6249,

Universite de Franche-Comte, Besancon, France

B. Quesnel � P. Boulvais

Geosciences Rennes, CNRS, UMR 6118,

Universite de Rennes 1, Rennes, France

C. Couteau

Service Geologique, Koniambo Nickel SAS,

Nouvelle Caledonie, France

123

Contrib Mineral Petrol (2014) 167:952

DOI 10.1007/s00410-013-0952-8

Page 59: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

et al. 2009; Rudge et al. 2010; Kelemen et al. 2011; Klein

and Garrido 2011). Among the various mineral species that

may undergo carbonation reaction, Mg-bearing minerals

(as well as Ca-carbonates) are of a great interest as they are

very common at Earth surface and thus represent an

important reservoir for CO2. In addition, magnesite

(MgCO3) has a long-term stability, contrarily to alkali

carbonates, which are readily soluble in water (Lackner

et al. 1995). Basically, magnesite is formed by the reaction

between a Mg-rich source and CO2-rich fluids (Bain 1924)

and can integrate variable amount of cations (mainly

divalent, e.g., Ca and Fe) by substituting Mg in the crystal

structure. Based on the study of numerous magnesite

deposits, Abu-Jaber and Kimberley (1992b) distinguished

vein-type and massive-type magnesite, the second type

forming deeper than the former. Numerous other parame-

ters play a role in magnesite formation, such as tempera-

ture, the origin of the components or the mechanisms of

precipitation. The temperature of magnesite formation

extends from ambient to *400 �C (Halls and Zhao 1995;

Wilson et al. 2009; Klein and Garrido 2011). The origin of

CO2 is variable as it may be related to either weathering

(atmospheric CO2), metamorphic (deep seated CO2) or

magmatic (magmatic CO2) sources (Abu-Jaber and Kim-

berley 1992b). The source of magnesium is usually local at

the outcrop scale but can also be distant, e.g., coming from

the weathering of magnesian rocks at the Earth surface and

transported downward by meteoric water infiltration

(Podwojewski 1995; Jurkovic et al. 2012). Mechanisms of

magnesite precipitation are also variable. Abu-Jaber and

Kimberley (1992b) reported two main ways of magnesite

precipitation: (1) precipitation through a reaction that

involves CO2-rich fluid and Mg-rich rock, or alternatively

Mg-rich fluids and CO2-rich rock (i.e., the most common

way to form magnesite); (2) the oversaturation of the fluid

with respect to magnesite may be enhanced by the fluid

evaporation and/or degassing (Dabitzias 1980; Fallick et al.

1991; Zedef et al. 2000; Ghoneim et al. 2003). In both

cases, hydrated Mg-carbonate species may precipitate

alternatively or in association with magnesite (e.g., Zedef

et al. 2000; Beinlich and Austrheim 2012) Detailed studies

of the precipitation mechanisms can thus provide major

clues for the understanding of natural CO2 sequestration.

Among the various rocks that have the potential to react

with CO2, ultramafic rocks, and particularly serpentinites,

are probably the most efficient feedstock material for long-

time storage through the formation of magnesite (Dabitzias

1980; Jedrysek and Halas 1990; Pohl 1990; Fallick et al.

1991; Abu-Jaber and Kimberley 1992b; Sherlock and

Logan 1995; Goff and Lackner 1998; Gerdemann et al.

2003; Ghoneim et al. 2003; Cipolli et al. 2004; Schulze

et al. 2004; Hansen et al. 2005; Teir et al. 2007, 2009;

Kelemen and Matter 2008; Rudge et al. 2010; Klein and

Garrido 2011; Jurkovic et al. 2012). Serpentine carbonation

onsets by the dissolution of atmospheric CO2 into water,

where CO2 forms different species as a function of pH. At

pH \ 6.5, carbonic acid (H2CO3) dominates, at pH

between 6.5 and 10.5, bicarbonate (HCO3-) dominates and

at higher pH, carbonate anion (CO32-) dominates. Ser-

pentine dissolution and magnesite precipitation are also

pH-dependent (Klein and Langmuir 1987; Guthrie et al.

2001; Teir et al. 2007; Prigiobbe et al. 2009; Teir et al.

2009; Krevor and Lackner 2011). Experimentally, Teir

et al. (2007) have shown that the best efficiency for ser-

pentine to carbonate conversion is obtained in the pH range

of 8–11, with an optimum at pH 9. Roughly similar pH

conditions were measured for the optimum carbonation of

olivine (Prigiobbe et al. 2009). At pH [ 8, bicarbonate

starts to dissociate into H? and CO32- ions:

HCO�3 ! Hþ þ CO2�

3 ð1ÞInteraction between CO2-rich water and serpentine gives

rise to the exchange of H? and Mg2? cations on the min-

eral surface. This reaction produces silica and water, while

free Mg2? cations react with CO32- anions to form

magnesite:

Mg3Si2O5 OHð Þ4þ 6Hþ ! 3Mg2þ þ 2H4SiO4 þ H2O

ð2ÞMg2þ þ CO2�

3 ! MgCO3 ð3ÞThe overall reaction can be summarized as follow:

Mg3Si2O5 OHð Þ4Serpentine

þ 3 2Hþ þ CO2�3

� �

CO2ðaqÞ

! 3MgCO3Magnesite

þ 2SiO2Silica aqð Þ

þ 5H2OWater

ð4Þ

Worldwide, carbonation of serpentinite commonly

leads to the formation of magnesite deposits in associ-

ation with ophiolitic bodies, e.g., in California (Bodenlos

1950; Sherlock and Logan 1995), Egypt (Ghoneim et al.

2003), Greece (Dabitzias 1980), Italy (Cipolli et al.

2004), Norway (Beinlich et al. 2012) or Oman (Kelemen

and Matter 2008). In New Caledonia, similar magnesite

deposits have been described in the serpentine sole of

the ophiolite, but few studies have focused on the origin

of this magnesite (Glasser 1904; Trescases 1973; Ulrich

2010; Quesnel et al. 2013). Quesnel et al. (2013) propose

that magnesite veins in the serpentinite sole formed

tectonically, at temperatures between 40 and 80 �C. On

the basis of stable isotope analyses (O and C), these

authors propose that the fluid from which magnesite

formed was originally meteoric water and that the car-

bon source is mostly atmospheric, with a possible bio-

genic contribution. They also suggest that magnesium

originated from the dissolution of the peridotite at the

952 Page 2 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 60: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

top of the ophiolite during laterization and was trans-

ferred down to the sole by the infiltration of meteoric

waters under tectonically active conditions. The per

descensum model is largely invoked to explain magne-

site deposits in lateritic environments (Dabitzias 1980;

Pefrov et al. 1980; Pohl 1990; Abu-Jaber and Kimberley

1992a, b; Foster and Eggleton 2002; Jurkovic et al.

2012; Oskierski et al. 2012). It requires a well-drained

system where meteoric waters, charged with atmospheric

CO2, dissolve serpentine (and other Mg-silicates), a

process that releases magnesium and silicon into solu-

tion. Meteoric waters then percolate downward thanks to

a microfracturing permeability system, thus dissolving

more magnesium. Within this frame, precipitation of

magnesite is due to supersaturation of fluids, which are

generated by the neutralization of carbonic acid by ser-

pentine dissolution, as shown by reactions (2) and (3)

(Bodenlos 1950; Pohl 1990; Fallick et al. 1991; Abu-

Jaber and Kimberley 1992b; Giammar et al. 2005; Kel-

emen et al. 2011). Alternatively, the ultramafic pile may

become a partially closed system, and therefore limit the

exchange between atmospheric CO2 and meteoric water

(Jurkovic et al. 2012). Progressive dissolution of ser-

pentine during the fluid migration downwards consumes

H? ions, as shown by reaction (3). As CO2 is not freely

available anymore, such consumption of H? ions causes

the increase in pH, enhancing magnesite precipitation

(Jurkovic et al. 2012).

Considering the origin of magnesium as deriving from

the laterization rises the question about the mechanism of

magnesite precipitation in New Caledonia. In such a case,

and similarly to magnesite deposits from Euboea (Greece,

Boydell 1921), magnesite formation may be the result of

direct precipitation from the fluids without interacting

with hosted rocks of the deposit site. This hypothesis was

favored by the composition of New Caledonia waters

passing through peridotites and serpentinites (sampled in

la Coulee river, Mont-Dore), showing that they are

enriched in magnesium and CO32- ions (Barnes et al.

1978). Alternatively, field investigations highlight a close

relationship between serpentinite, magnesite and silica,

suggesting that the two latter phases may originate from

the dissolution of the former. In this case, magnesite

would precipitate by in situ replacement of the serpen-

tine, using magnesium released during the serpentine

dissolution. In this study, we provide evidences of mag-

nesite formation through such a process on the basis of

mineral characterization using an original analytical

method for the interpretation of micro-X-ray fluorescence

(l-XRF) images. We particularly focus on the formation

process of magnesite veins coupled to intense silicifica-

tion observed in the serpentinite sole of the New Cale-

donia ophiolite.

Geological settings and sample descriptions

New Caledonia is located in the SW Pacific, 2,000 km east

of the Australian coasts (Fig. 1). It is composed of several

islands that belong to the Norfolk ridge (La Grande Terre,

Island of Pines, Belep Islands) and to the Loyalty ridge

(Loyalty Islands). The main island, la Grande Terre, con-

sists of a patchwork of terranes reflecting the geodynamic

evolution of the SW Pacific region from late Permien to

Eocene (e.g., Cluzel and Meffre 2002; Cluzel et al. 2001,

2012). Among these terranes, the ophiolite is the most

prominent as it covers more than 25 % of the island. The

so-called peridotite nappe is composed of a large and

continuous massif located south of the island and some

isolated klippes widespread along the west coast (Fig. 1).

The ophiolite was formed between 83 Ma with the opening

of the South Loyalty Basin and its subsequent closing at

34 Ma, timing of its obduction on the Norfolk continental

basement (Cluzel et al. 2001; Crawford et al. 2003;

Schellart et al. 2006; Whattam et al. 2008; Whattam 2009;

Ulrich et al. 2010; Cluzel et al. 2012). Since its emergence,

the uppermost part of the peridotite nappe has undergone

an intense laterization. This led to the development of a

thick laterite bed (up to 60 m thick; Sevin et al. 2012) that

owns *30 % of the world nickel resources. The whole

ophiolite is formed of upper mantle rocks (mainly harz-

burgites) with minor cumulates (Prinzhoffer 1981). Peri-

dotites are highly serpentinized, particularly at the base of

the ophiolite which is made of a thick (up to 400 m thick,

Audet 2008; Ulrich et al. 2010) and silicified serpentinite

sole where large amount of magnesite veins have crystal-

lized (Quesnel et al. 2013).

Samples presented in this study were collected in the

serpentinite sole of the Koniambo massif (Fig. 1). Similar

outcrops occur on the serpentinite sole from other perido-

tite massifs in New Caledonia (Ulrich 2010). In the field,

the whole sole is highly deformed, finely schistose and/or

intensely brecciated, and has recorded multiple serpenti-

nization events (Ulrich 2010): Massive serpentinization,

which is of a typical bottle green color in the field (Fig. 2),

is cross-cutted by light-green-colored serpentine veins. The

latter is associated with black magnetite impregnations and

microcracks filled by fibrous serpentine (chrysotile).

Magnesite occurs as millimeter to multi-decimeter stock-

work veins with a typical cauliflower-like texture, cross-

cutting serpentinites (Fig. 2). The magnesite is mainly

snow white colored, but can also appear greenish depend-

ing of the amount of intergrown serpentine. On the basis of

structural observations, Quesnel et al. (2013) distinguished

two types of veins. The first type of veins is observed along

and/or within the margins of top to the SW-centimeter

shallow dipping shear zones. The second type of veins

corresponds to steeper veins occasionally crosscut by the

Contrib Mineral Petrol (2014) 167:952 Page 3 of 19 952

123

Page 61: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

low-dipping shear zones. Magnesite can also develop per-

vasively by precipitating in large and massive serpentine

blocks (Fig. 2a–c). In this case, the magnesite seems to

develop first at the expense of the light-green-colored

serpentine to progressively extend to bottle green serpen-

tine, a relationship that is not clear when magnesite fills the

main shear zones, as described by Quesnel et al. (2013).

Analytical methods

X-ray diffraction

Analyses were performed at the Institut des Sciences de la

Terre (ISTerre, Grenoble, France) on sample powders

obtained after the crushing of separated magnesite veins

and serpentinite host rock mineral fractions. X-ray dif-

fraction (XRD) patterns were recorded with a Bruker

D5000 powder diffractometer equipped with a Kevex

Si(Li) solid state detector using CuKa1?2 radiation.

Intensities were recorded at 0.02� 2h step intervals from 5

to 80�, with a 6-s counting time per step. Size of the

divergence slit was 0.298�.

Raman spectroscopy

Raman spectroscopy measurements were performed at the

Ecole Normale Superieure of Lyon and at GeoRessources

Nancy, France, in both cases using a Horiba Jobin–Yvon

LabRam HR800 spectrometer and a visible ionized argon

laser source with a wavelength of 514 nm. Output laser

power was 100 mW, and measurements were performed

using an Olympus lens of 9100 to focus the laser beam

onto an area that was 1 lm in diameter. Analyses were

carried out on macroscopic samples and on thin sections.

Spectra result from the average of 5 acquisitions of 10–20 s

to optimize the signal/noise ratio. Two regions of the

Raman spectra were investigated: 150–1,250 cm-1 for

structural bonding characterization and 2,800–3,900 cm-1

for the characterization of the hydroxyl groups.

Micro-X-ray fluorescence

Micro-X-ray fluorescence analyses were performed on a

5-mm-thick rock section (Fig. 2d) using EDAX Eagle III

spectrometer at ISTerre (Grenoble, France). The X-ray

tube consists of a Rh anode operating at 250 lA with an

acceleration voltage of 40 kV. Polycapillary lenses were

used to focus the X-ray beam down to 40 lm full-width-at-

half-maximum at the sample surface. An energy-dispersive

X-ray detector with resolution of 140 eV was used to

measure fluorescence spectra. Chemical maps were recor-

ded with a matrix of 256 9 200 pixels, a 40-lm step

interval in both directions, and a dwell time of 1 s per

pixel. For each map, the gray scale corresponds to the

intensity of the Ka-lines of the different elements (Si, Mg,

50 km

Peridotite nappe

Touho

Poindimié

Ile Ouen

Koumac

Voh

Pouébo

Belep Islands

Nouméa

Bourail

Poya

Kunié

Plum

Thio

Tiebaghi

Massif du Sud

Kopéto

TASMAN SEA New Zealand

Vanuatu Arc

Fiji

Australia

NewCaledonia

PACIFICOCEAN

CORAL SEA

Sampled peridotite massif

166°E164°E

20°S

22°S

168°E

168°E166°E164°E

20°S

22°SMain lateritic beds

5 km

N

Koniambo

Serpentine sole

Undifferenciated basements

Loyalty Islands

Norfolk ridge

Loyalty ridge

Lord Howeridge

Fig. 1 Localization of the main ophiolitic occurrences composing the peridotite nappe. Simplified geological map of the Koniambo massif is

modified from Maurizot et al. (2002)

952 Page 4 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 62: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Fe, Al, Ni, Ti, Mn, Cr, Ca, K) calculated from the inte-

gration of a specified region of interest (ROI) of the energy

range of XRF spectra. Then, ROI maps (see electronic

supplements, Figure S4) are used to calculate phase maps

thanks to a new routine, specially developed Matlab�-

based code, following the same approach than that suc-

cessfully applied first to the computing of mineral-phase

maps from hyperspectral l-XANES mapping (Munoz et al.

Fig. 2 a, b, c Snow-white-colored-magnesite veins crystallized in

massive serpentinite blocs, Koniambo massif. Magnesite shows a

typical cauliflower texture and is closely associated with silica veins

(in brown). d Typical sample of carbonated serpentinite. Orange

dashed line square localizes the mapped area by l-XRF presented in

Figs. 6, 7, 8 and in Figures S2 and S3

Table 1 In situ concentration measurements by l-XRF for serpentine, magnesite, silica, sepiolite, magnetite and chromite used for the

calculation of quantitative maps shown in Fig. 6 (see also electronic supplements, Figure S6)

Elements Serpentine Magnesite Silica Magnetite Sepiolite Chromite

MgO 40.25 46.84 – – 24.72 12.34

Al2O3 0.41 – – – 0.46 23.22

SiO2 43.60 0.39 100.00 – 62.66 0.08

K2O – – – – 0.01 0.01

CaO – 0.54 – – 0.2 0.07

TiO2 0.01 – – – 0.02 –

Cr2O3 0.02 – – – 0.02 48.20

MnO 0.03 – – – 0.67 0.20

FeO 2.51 0.02 – 100.00 0.55 15.82

NiO 0.17 – – – 0.11 0.06

CO2 – 52.21 – – – –

H2O 13.00 – – – 10.58 –

Total 100.00 100.00 100.00 100.00 100.00 100.00

Contrib Mineral Petrol (2014) 167:952 Page 5 of 19 952

123

Page 63: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

2008). In the case of hyperspectral l-XRF maps, the phase

map calculation consists first of determining pure mineral

phases that are expected to be present in the sample, in

order to create standard spectra (or ‘‘pure’’ spectra). Then,

for each pixel of the map, a linear combination of the

different standard spectra is performed in order to fit each

single spectrum. Results provide quantitative phase maps

showing the distribution of minerals previously identified

in the sample (e.g., based on XRD and/or Raman analyses).

This approach is particularly useful to highlight relation-

ships between minerals, especially for the characterization

of finely divided mineral assemblages, i.e., when the beam

is larger than grain size (such as here, typically below 1

micron). Concentration maps (in wt%) are finally recal-

culated on the basis of phase distribution maps considering

the chemical composition of standards (Table 1).

Results

Mineralogy and chemical compositions

Bulk mineralogical compositions

Figure 3 shows the XRD patterns obtained on separated

fractions of serpentinite and magnesite. The serpentinite

mainly consists of serpentine with minor amount of

magnetite and chromite. Serpentine occurs as lizardite

and chrysotile. The occurrence of chrysotile is consistent

with the presence of fibrous serpentine in microcracks as

described above. Despite a careful separation, small

peaks of magnesite and silica are present in the dif-

fraction pattern of the serpentinite, suggesting that both

minerals also occur at the micrometer scale within the

hosted rock.

X-ray diffraction pattern of magnesite powder shows

that the carbonate exhibits its most characteristic reflec-

tions at the following d values (in A, arranged according

to decreasing intensities): 2.74, 2.10, 1.70, 2.50 and 1.94

(Fig. 3). In addition, the diffraction pattern shows that

magnesite systematically integrates small amounts of

sepiolite (Mg4Si6O15(OH)2�6(H2O)), a mineral that is

frequently described in association with carbonate in

ultramafic environments (e.g., Birsoy 2002; Yalicin and

Bozkaya 2004; Boschetti and Toscani 2008). On the

basis of this XRD pattern, magnesite is the only car-

bonate to crystallize. Neither dolomite nor calcite are

formed, contrarily to numerous magnesite deposits pre-

viously described in the literature (Griffis 1972; Dabitz-

ias 1980; Jedrysek and Halas 1990; Fallick et al. 1991;

Abu-Jaber and Kimberley 1992b; Lugli et al. 2000; Ze-

def et al. 2000; Ghoneim et al. 2003; Jurkovic et al.

2012).

Optical microscopy

Figure 4 shows the typical mineralogical textures of car-

bonated serpentinites composing the sole of the New Cal-

edonia ophiolite. In thin section, the serpentinite does not

exhibit any relic of primary minerals (i.e., olivine and

pyroxene). However, the habits of grains and original

textures of primary minerals (e.g., cleavage planes of

orthopyroxene) have been preserved (typical pseudomor-

phic ‘‘mesh’’ texture, Fig. 4a, b) and indicate that the

parent rock was a harzburgite. As highlighted by XRD

analyses, opaque minerals associated with the serpentine

are small grains of magnetite and chromite disseminated in

Liz

Chr

Liz

Opl

Liz

Chr

Mst

Mgt

Ctl

Ctl

Liz

CtlLi

zCtl

Liz M

st

Chr

Liz

Liz

Chr

CtlLiz

Ctl

Liz

0

1

2

3

4

5

6

7

8

2 θ Scale10 20 30 40 50 60 70

Counts (x10 )

80

Serpentinite

0

2

4

6

8

10

12

14

16

18

20

22

24

26

28

30

32

2 θ Scale10 20 30 40 50 60 70

Counts (x10 )

80

Sepiolite

Magnesite

2.74

2.10

1.70

2.501.94

1.34

1.35

2.31 1.77

1.51

1.49

1.41

1.37

1.25

1.24

Fig. 3 Typical XRD patterns of serpentinite and magnesite. Liz

lizardite, Chr chromite, Opl opal, Mgs magnesite, Mgt magnetite and

Ctl chrysotile. Numbers upon the magnesite peaks correspond to

d values, given in A

952 Page 6 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 64: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the mesh. In these rocks, magnesite occurs as finely dis-

seminate grains (i.e., cryptocrystalline texture) that devel-

oped both at the rim and in cracks affecting serpentine

grains (Fig. 4b, d). These grains progressively aggregate to

form larger zones with a typical granular texture (Fig. 4a–

c). The sepiolite, identified by XRD and optical micros-

copy, occurs as brown fibers interstitially to the magnesite

nodules (Fig. 4c). Silica also occurs close to magnesite.

Microscopic observations show that the nature of the silica

is variable: it occurs as an amorphous solid (gel-like)

Fig. 4 a Microphotography under polarized light illustrating the

development of magnesite and silica in serpentinite. b Microphotog-

raphy under crossed-polarized light showing the development of

magnesite along the rims of serpentine grains (indicated by the red

arrows). Notice that the serpentinization also affects the orthopyrox-

ene. c Magnesite grain aggregates surrounded by colloform amor-

phous silica gel. Associated brown and fibrous mineral corresponds to

sepiolite (observation under polarized light). d Nucleation of

magnesite grain on serpentine surface surrounded by colloform

amorphous silica gel. Black dots inside of the serpentine grain

correspond to magnetite (observation under crossed-polarized light).

e Microtextures of a silica vein under polarized light (mesh:

serpentine mesh; #1: opal-CT, #2 and #3: Chalcedony; identifications

made by raman spectroscopy, see Fig. 5). f Same as e under crossed-

polarized light. Mineral abbreviations: Mgs magnesite, Mgt magne-

tite, Opx orthopyroxene, Sep sepiolite and Srp serpentine

Contrib Mineral Petrol (2014) 167:952 Page 7 of 19 952

123

Page 65: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

surrounding magnesite aggregates and serpentine grains in

area where magnesite dominates (Fig. 4a, c, d). Where

serpentine is dominant, silica consists of crystalline to fine

grains, forming vugs with a typical colloform texture

propagating in the serpentine mesh (Fig. 4e, f). In this case,

the nature of silica ranges from amorphous-like near the

rim of the vug to quartz-like at the center (Fig. 4c, e, f).

Raman spectroscopy

Raman analyses were performed to further identify min-

erals that compose the serpentinite. This technique is

complementary to XRD, since it is particularly efficient to

distinguish among the different varieties of serpentine (e.g.,

Lemaire 2000; Auzende et al. 2004) as well as silica

polymorphs (e.g., Gotze et al. 1998; Pop et al. 2004).

Figure 5 shows Raman spectra collected on carbonated

serpentinites. Results show that lizardite is the dominant

serpentine variety that occurs as individual grains and in

the mesh, and corresponds to the bottle-green-colored

serpentine described in the macroscopic observations.

Chrysotile (not shown in Fig. 5) has also been detected, as

already highlighted by XRD. In addition, light green ser-

pentine was identified as polygonal serpentine. It shows

quite similar patterns to those of the lizardite at low wave

numbers, but strongly differs at high wave numbers (from

3,500 to 3,800 cm-1, corresponding to OH group), where

the polygonal serpentine is characterized by a large peak

composed by two bands centered at 3,689 and 3,700 cm-1

(e.g., Lemaire 2000; Auzende et al. 2004).

The Raman spectrum of magnesite is characterized by

four distinct bands located at 209, 327, 737 and

1,094 cm-1 (Fig. 5), consistently with the work of Krish-

namurti (1956). The lack of bands at 3,448 and 3,648 cm-1

(i.e., in the OH region; typical of hydromagnesite) shows

that magnesite is anhydrous.

Silica polymorphs are clearly identified using Raman

spectroscopy. Vug rims consist of opal-CT (Silica #1,

Fig. 4e), while brown coronas (Silica #2, Fig. 4e) and

white fine grains (Silica #3, Fig. 4e) inside of the vugs are

identified as chalcedony (Fig. 5). Although significantly

different in microscopic observations (Fig. 4e, f), both

chalcedonies display very similar Raman spectra, except on

the intensity of the band located at 501 cm-1 which is

significantly higher in the white chalcedony (electronic

supplements, Figure S1). Notice that in all silica poly-

morphs, the large bands observed between 3,100 and

3,900 cm-1 indicate the presence of molecular water.

XRF mapping

The elemental distribution, expressed in weight percent, in

a typical serpentinite texture surrounded by magnesite is

shown in Fig. 6. On the basis of chemical measurements, it

was not possible to discriminate lizardite from polygonal

serpentine, both having very similar compositions. There-

fore, we only refer to serpentine in the following, focusing

on the nature of processes that preferentially affect the

serpentine grains or the mesh. The distribution maps of

MgO and SiO2 highlight serpentine grains (s) as well as

mesh texture (m). The chemical composition of these

grains is consistent with the stoichiometry of serpentine

minerals, with about 43 wt% for both MgO ? Fe2O3 and

SiO2. However, SiO2 concentration in the mesh is signifi-

cantly higher and can reach up to 65 wt%. To better

understand such differences in chemical compositions, we

calculated the phase distribution maps according to the

minerals that are expected to be present in this sample

(Fig. 7 and electronic supplements, Figure S4). Phase dis-

tribution reveals that the central zone of the mapped area

mainly consists of serpentine minerals. The amount of

serpentine is at least 50 % in the mesh and close to 100 %

209

327

737

1094

201

465

501

801

232

351

390

620

690

350

3688

3706

3654

Inte

nsity

(a.

u)

3314

3429

3575

33143429

3575

200 400 600 800

Wavenumber (cm-1)

1000 1200

3650 3700 3750

3200 3400 3600 3800

3200 3400 3600 3800

226301

Lizardite

Polygonal

3689 3700

3648

231

346

389

620

690S

erpe

ntin

eC

halc

edon

yO

pal C

TM

agne

site

Fig. 5 Typical Raman spectra observed for serpentine (lizardite and

polygonal), chalcedony, opal-CT and magnesite composing the

serpentinite sole of the New Caledonia ophiolite

952 Page 8 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 66: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

in grains. Silica distribution is relatively homogeneous in

the mesh, with about 30 % silica, but shows peculiarities

close to the magnesite, where the silica content reaches

about 100 %. In contrast, the amount of silica in serpentine

grains is well below 5 %. Finally, the calculation of phase

distribution maps was obtained with about 20 % of mag-

nesite in the serpentinite texture (i.e., the central area).

Figure 8 shows density correlation diagrams between

serpentine, magnesite and silica based on the phase dis-

tribution maps for the two regions delimited by squares in

Fig. 7. These diagrams statistically illustrate the descrip-

tion made above on the phase maps and reveal two distinct

processes. In the region 1, the diagrams show a clear an-

ticorrelation between silica and serpentine, corroborating

that a great part of the pixels of this region corresponds to

‘‘silicified serpentine.’’ At the opposite, the relationships

between magnesite and serpentine on one hand and those

between silica and magnesite on the other hand are less

straightforward. To better understand these two diagrams,

the region 2 (dashed square in Fig. 7) was delimited around

a heterogeneous serpentine grain that appears partially

altered (Fig. 9). This alteration occurs around fractures in

the grain and is expressed by a lighter color of the ser-

pentine in microscopic view. Considering this grain, a

statistical analysis shows that silica and magnesite are

correlated with each other, whereas they are both anticor-

related with serpentine (Fig. 8, region 2). This point is

particularly interesting since the phase maps clearly high-

light here a process of replacement of serpentine by an

assemblage of magnesite and silica. Even if silica and

serpentine are anticorrelated in both regions, carbonation

and silicification processes do not follow rigorously the

same trend. On the basis of our specific treatment of the

XRF maps, we demonstrate here that the serpentinite is

affected by two distinct weathering processes: while the

mesh texture is mainly affected by a silicification process,

the serpentine grains are mainly affected by the crystalli-

zation of a magnesite ? amorphous silica assemblage.

Discussion

Serpentinite carbonation in New Caledonia has a supergene

origin, following a per descensum model (Glasser 1904;

Trescases 1973; Ulrich 2010; Quesnel et al. 2013). The

atmospheric carbon dioxide is first dissolved in meteoric

water. The CO2-enriched fluids then circulate through the

lateritic cover and dissolve the residual Mg-rich minerals

Optical

mm

mm

0 5 10 15

0

2

4

6

8

10

mm

mm

MgO (wt.%)

0 5 10 15

0

2

4

6

8

10

mm

mm

0 5 10 15

0

2

4

6

8

10

mm

mm

FeO (wt.%)SiO2 (wt.%)

0 5 10 15

0

2

4

6

8

10

0

10

20

30

40

50

60

70

80

90

100

0

5

10

15

20

25

30

35

40

45

50

55

0

10

20

30

40

50

60

70

80

90

100

Srp

mesh

Mgs

Srp

m

Mgs

Srp

mesh

Mgs

Srp

mesh

Mgs

Fig. 6 Quantitative chemical maps of MgO, SiO2 and Fe2O3 (in wt%) calculated on the basis of l-XRF measurements (EDAX Eagle III).

Complete procedure for the map calculation is detailed in the text

Contrib Mineral Petrol (2014) 167:952 Page 9 of 19 952

123

Page 67: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

before driving to the precipitation of magnesite in fractures

and to a porosity reduction in the serpentinite, within the

serpentinite sole. This carbonation process leads to the

formation of clusters of magnesite, potentially very large,

and usually very localized to the serpentine-rich fractures

or volumes. Although this process is most easily observed

in the field, our results show that the serpentine dissolution

leads to the local crystallization of magnesite and silica.

This diffuse process is potentially the main mechanism of

carbonation of the serpentinite sole.

Carbonation after serpentine dissolution

On the basis of our results of our mineralogical investi-

gations, several features provide evidence that magnesite

precipitation occurs through serpentine dissolution. First,

the development of magnesite along the edges of adjacent

serpentine grains is a characteristic of carbonation by

in situ replacement of the serpentine (Figs. 4b, d, 9). When

serpentine grains are preserved, carbonation is limited to

the rims, the core remaining unaffected. At the opposite,

the dissolution and replacement of the serpentine are more

efficient in case of fractured grains (Fig. 9). This obser-

vation is consistent with experimental studies showing that

intense grain fracturing is required to ensure complete

carbonation (e.g., Haug et al. 2011; Kelemen et al. 2011;

Hovelmann et al. 2012; van Noort et al. 2013). Another

way to illustrate the progressive serpentine carbonation is

given by Figure S2 (in electronic supplements). Here, the

RGB map (for Red, Green, Blue, see the figure caption for

more details) shows that magnesite is ubiquitous, formed

by the replacement of the serpentine grains, and that only a

few of the latter are preserved from the carbonation pro-

cess. In addition, considering the stoichiometry and the

molar volumes of the mineral species involved in reaction

(4), volumes of magnesite and silica produced by the dis-

solution of 1 mol of serpentine (108 cm3) are, respectively,

85 and 58 cm3 (VMgs/VSilica * 1.45), leading to a volume

increase in about 30 %. Statistics calculated on the ser-

pentine grain presented in Fig. 9 (corresponding to the

region 2, Fig. 7) show that the assemblage consists of about

62 vol% of serpentine, 21 vol.% of magnesite and

15 vol.% of silica (VMgs/VSilica * 1.4; Fig. 8). This

assemblage compares well with the result of dissolution of

38 vol.% of serpentine, which leads to the formation of

*22 vol.% of magnesite and *15 vol.% of silica. These

estimates testify that results given in the phase maps are in

good agreement with the stoichiometry of reaction (4).

mm

mm

Magnesite (%)

0 5 10 15

0

2

4

6

8

10

0

10

20

30

40

50

60

70

80

90

100

Optical

mm

mm

0 5 10 15

0

2

4

6

8

10

mm

mm

Serpentine (%)

0 5 10 15

0

2

4

6

8

10

0

10

20

30

40

50

60

70

80

90

100

mm

mm

Silica (%)

0 5 10 15

0

2

4

6

8

10

0

10

20

30

40

50

60

70

80

90

100

1

2

Fig. 7 Spatial repartition and quantification of mineral phases (in %) calculated on each pixel of the mapped area (Fig. 2d), based on l-XRF

measurements (EDAX Eagle III). Square-delimited regions correspond to areas used to construct correlation diagrams in Fig. 8

952 Page 10 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 68: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Second, carbonation by serpentine dissolution is also evi-

denced by the precipitation of an amorphous silica layer at

the grain rims (Fig. 4d). This observation is consistent with

the previous studies that systematically reported the for-

mation of a Si-rich layer after Mg-bearing mineral break-

down (e.g., Luce et al. 1972; Lin and Clemency 1981;

Serpentine (%)

Sili

ca (

%)

20 40 60 80 100

20

40

60

80

100

Serpentine (%)

Mag

nesi

te (

%)

20 40 60 80 100

20

40

60

80

100

Magnesite (%)

Sili

ca (

%)

20 40 60 80 100

20

40

60

80

100

Serpentine (%)

Sili

ca (

%)

20 40 60 80 100

20

40

60

80

100

Serpentine (%)

Mag

nesi

te (

%)

20 40 60 80 100

20

40

60

80

100

Magnesite (%)

Sili

ca (

%)

20 40 60 80 100

20

40

60

80

100

0

100

200

300

400

500

600

700

800

900

1000

0

100

200

300

400

500

600

0

100

200

300

400

500

600

700

510152025303540455055

0

5

10

15

20

25

30

35

40

45

10

20

30

40

50

60

Region 1 (n = 9200)

Region 2 (n = 256)

0

Serpentine carbonation

Mesh silicification

Fig. 8 Correlation diagrams (represented as density fields) between

serpentine, magnesite and silica based on the phase distribution maps

for the two square-delimited regions in Fig. 8. n corresponds to the

total number of pixels composing each region. Colorbar values

correspond to the number of pixels in one grid cell of a size of 5 % by

5 %

Chromite

Magnetite

Serpentine

200 µm

Fig. 9 Microphotography under polarized light and schematic representation of the partially carbonated serpentine grain (corresponding to the

Region 2 in Fig. 7)

Contrib Mineral Petrol (2014) 167:952 Page 11 of 19 952

123

Page 69: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Guthrie et al. 2001; Schulze et al. 2004; Giammar et al.

2005; Bearat et al. 2006; Sipila et al. 2008; Andreani et al.

2009; Daval et al. 2011; King et al. 2011; Hovelmann et al.

2011, 2012). Third, sepiolite, which is described in our

samples in association with magnesite (Figs. 3, 4c), also

formed subsequently to the dissolution of serpentine (e.g.,

Jones and Galan 1988; Birsoy 2002; Yalicin and Bozkaya

2004; Andreani et al. 2009). Similar occurrences of sepi-

olite have been reported during natural (Yalicin and Boz-

kaya 2004) and experimental (Andreani et al. 2009)

carbonation of Mg-bearing minerals (serpentine and oliv-

ine). Actually, sepiolite may form in different ways:

2Mg3Si2O5 OHð Þ4Serpentine

þ 2 2Hþ þ CO2�3

� �

CO2ðaqÞ

þ 2H4SiO4SilicaðaqÞ

! 2MgCO3Magnesite

þMg4Si6O15 OHð Þ2�6H2OSepiolite

þ 3H2OWater

ð5Þ

3Mg3Si2O5 OHð Þ4Serpentine

þ 5 2Hþ þ CO2�3

� �

CO2ðaqÞ

! 5MgCO3Magnesite

þMg4Si6O15 OHð Þ2�6H2OSepiolite

þ 4H2OWater

ð6Þ

Mg3Si2O5 OHð Þ4Serpentine

þ 2.5 2Hþ þ CO2�3

� �

CO2ðaqÞ

! 2.5MgCO3Magnesite

þ 0.125Mg4Si6O15 OHð Þ2�6H2OSepiolite

þ 1:25H4SiO4Silica aqð Þ

þ 1.125H2OWater

ð7Þ

All these reactions contribute to form magnesite, but

they differ as reactions (5) and (6) conserve all Mg and Si

in solids and are characterized by a volume gain (*60 and

*30 %, respectively) while reaction (7) is balanced on

volume. In addition, Eq. (5) consumes silica while Eq. (7)

produces it. Only on the basis of cm-scale maps, it is dif-

ficult to estimate which reaction occured: in our sample,

sepiolite represents \1 vol.% (see electronic supplements,

Figure S4). Among the reactions written above, Eq. (7) is

the one that produces the smallest amount of sepiolite,

regarding the amount of magnesite that stoichiometrically

precipitates. In contrast, Birsoy (2002) demonstrated that

sepiolite formation is much more favored in the presence of

Si-rich solution. In that case, sepiolite in our sample may

derive from Eq. (5).

The fact that some serpentine grains are only partially

transformed into magnesite indicates that the carbonation

process was not completed. The reason why the reaction

does not go to completion may be due to (1) silica pre-

cipitation, (2) pH increase and/or (3) porosity decrease.

Numerous studies proposed that the formation of a Si-rich

layer at the rims of serpentine grain during the first steps

of dissolution might inhibit further Mg diffusion, poten-

tially retarding or even stopping the process of carbon-

ation (Gerdemann et al. 2003; Schulze et al. 2004;

Alexander et al. 2007; Daval et al. 2011). Although Si-

rich layer effectively reduces the accessibility of fluids to

the reactive surface of minerals, recent investigations

demonstrate that some permeability is maintained as car-

bonation remains active even after its development (Bea-

rat et al. 2006; Andreani et al. 2009; Hovelmann et al.

2012). Thus, it is not obvious that precipitation of Si-rich

layer led the serpentine carbonation to stop. Alternatively,

following reaction (4), the dissolution of serpentine con-

sumes 2 mol of H? for 1 mol of CO32-, leading to a

progressive increase in fluid pH. According to Barnes

et al. (1978), fluids reacting with serpentinite in New

Caledonia have pH ranging from 9 to 11. At pH 9, HCO3-

dominates over CO32-, making the serpentine dissolution

possible following reactions (1) and (2). At pH 10.5,

CO32- species become dominant and at pH 10.8, HCO3

-

represent \20 % of the carbonate species in the fluid. In

these conditions, magnesite precipitation is favored, but

serpentine dissolution is scarce due to the lack of H? ions

(e.g., Teir et al. 2007). Thus, intensive exchange between

H? ions and Mg2? cations on the serpentine surface leads

to a progressive increase in pH that may inhibit further

serpentine dissolution and subsequent carbonation. There

is no evidence to argue against such process in New

Caledonia, but it requires that atmospheric CO2 is not

freely available and thus a roughly closed system (Jurk-

ovic et al. 2012). A third explanation may account for the

inhibition of carbonation. Hovelmann et al. (2012)

recently investigated the microstructure and porosity

evolution as a function of carbonation reaction progress in

natural peridotite. On the basis of their experimental

results, they reported that a carbonation extend of *10 %

leads to a closure of 50 % of the initial porosity. They

demonstrated that magnesite precipitation in fracture pore

space reduces the permeability and progressively stops the

fluid pathway, preventing further reaction between the

fluid and the silicate surface and ultimately ends the car-

bonation process. Contrarily to serpentinization, which is

able to propagate through a reaction-induced fracturing

mechanism (e.g., Plumper et al. 2012), carbonation is self-

limited as the reaction will be inhibited due to magnesite

growing that ultimately clogs the system. This implies that

no volume gain occurs at the rock scale (Beinlich et al.

2012). Numerous studies highlight the necessity of active

fracturing to ensure complete carbonation (e.g., Kelemen

and Matter 2008; Kelemen et al. 2011). In our samples, no

evidence of carbonation-induced fracturing has been

observed so far, even if Quesnel et al. (2013) show the

syn-kinematic character of magnesite veins at the outcrop

scale. At the millimetric scale, volume gain associated

with reaction (4) may have inhibited the complete car-

bonation of serpentinite by clogging the reacting zone.

Such a process may prevent the infiltration of additional

fluids, resulting in the partial carbonation of serpentine

952 Page 12 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 70: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

grains as illustrated by Fig. 7 (see also electronic sup-

plements, Figure S2).

Redox conditions during magnesite precipitation

The behavior of iron during carbonation has been poorly

investigated. Such information would be indicative of

redox conditions during magnesite precipitation. In ser-

pentinite, iron is mainly hosted by iron oxides (magnetite,

Fe3O4) surrounding serpentine grains or disseminated in

the mesh texture (see electronic supplement, Figure S5).

Iron content in serpentine is *2 wt% (expressed as FeO,

Table 1). It substitutes Mg2? cations in octahedral sites

(O’Hanley 1996). During serpentine dissolution, iron is

released from the mineral structure and may subsequently

form Fe–Si-rich layer that progressively evolves into

magnetite at the interface between serpentine and aqueous

fluid (Fallick et al. 1991; Alexander et al. 2007; Andreani

et al. 2009; Saldi et al. 2013). Alternatively, it may pre-

cipitate as siderite (FeCO3) under rather high pCO2 and

specific pH range from 5.5 to 7.5 (Ohmoto et al. 2004) or

be integrated in magnesite by substituting Mg2? cations

(Abu-Jaber and Kimberley 1992b; Hansen et al. 2005). In

each of these minerals, iron oxidation is expected to differ

according to oxidizing conditions: it is mostly oxidized in

magnetite (i.e., 2/3 of iron is ferric iron) while it is in

ferrous state in siderite or magnesite. In our sample, XRF

analyses show that the iron content in magnesite is very

low (Table 1), suggesting that its integration into the

structure of the magnesite was very scarce. In addition, we

never identified siderite in our samples so far. As stated

above, siderite precipitates at pH conditions that are sig-

nificantly lower than those of New Caledonia waters, so it

is likely that pH was too high to make siderite precipitate.

Magnetite is ubiquitous in our sample. Therefore, dis-

criminating magnetite grains related to prior serpentiniza-

tion events to those potentially derived from carbonation is

not obvious. However, microscopic observations show that

numerous magnetite grains have precipitated inside of

partially carbonated serpentine grains (Figs. 4d, 9), while

magnetite related to serpentinization generally forms out-

side of serpentine grains. Schematic representation in

Fig. 9 highlights the systematic association of magnetite

with cracks in the serpentine grain. According to Andreani

et al. (2009), such grain fracturing may be considered as

zones of localized fluid flow that favor the precipitation of

magnetite during the first step of mineral dissolution. In

contrast, their experiments show that carbonation initiates

in domains of reduced fluid flow zones (fractures of a

smaller size) where chemical gradients are small and thus

facilitate local supersaturation, high pH and more reducing

conditions. These results indicate that magnetite associated

with the carbonation process may be used as a proxy to

estimate the fluid flow rate at the serpentine grain scale:

Fractures filled by magnetite correspond to high-fluid-flow

zones characterized by more oxidizing conditions. In these

regions, dissolution occurs but not magnesite precipitation

(Andreani et al. 2009). Magnesium migrates to zones of

reduced fluid flow, corresponding to regions dominated by

magnesite or partially carbonated serpentine (Figs. 9, 10).

In these areas, magnesite precipitates due to local super-

saturation under more reducing conditions. This underlies

that redox gradients occur even at grain scale.

Serpentine mesh silicification

Silica mobilization and precipitation

Intense silicification of the serpentine mesh was revealed

during this study based on XRF measurements and Raman

spectroscopy (Figs. 6, 7, 8, electronic supplements, Figures

S2 and S3). For instance, the region 1 (Fig. 7) is composed

of 62 vol.% of serpentine, 24 vol.% of silica and 7 vol.%

of magnesite in average. The precipitation of 7 vol.% of

magnesite consumes 12 vol.% of serpentine and forms

5 vol.% of silica, on the basis of reaction (4). Such excess

of silica necessarily involves a contribution of silica from

outside of the region 1. Phase maps show that pure silica

mainly occurs in gaps between magnesite aggregates (left-

side of the map, Fig. 7 and electronic supplements, Figure

S2). However, when aggregates agglomerate subsequently

to magnesite growth (right-side of the map, Fig. 7 and

electronic supplements, Figure S2), silica is expelled from

the magnesite. Numerous experimental studies indicate the

progressive migration of silica rather than being rejected

from the growing magnesite (Schulze et al. 2004; Hovel-

mann et al. 2011, 2012). As stated above, the first steps of

serpentine dissolution release silica that immediately pre-

cipitates as a Si-rich layer. As carbonation reaction pro-

ceeds, the released silicon may feed the growth of the Si-

rich layer, but Hovelmann et al. (2012) reported that such

inward growth is limited. This limitation is mainly asso-

ciated with pH increase in fluid during the carbonation

process. At pH [ 9, magnesite precipitates while silica is

solubilized and is therefore able to migrate away from the

solid–fluid interface. Occurrence of water within silica is

demonstrated by the Raman spectrum measured in the

region of high frequencies (Fig. 5), which exhibits a typical

spectrum of molecular water. As illustrated by XRF map-

ping, fluids enriched in silica subsequently to serpentine

dissolution propagate in the serpentine meshwork. Similar

observations have been reported in serpentinites from the

Oman and the Ligurian ophiolites (Stanger 1985; Boschi

et al. 2009; Lacinska and Styles 2012). According to

Lacinska and Styles (2012), the well-preserved mesh tex-

ture (as observed in our samples) induces a combination of

Contrib Mineral Petrol (2014) 167:952 Page 13 of 19 952

123

Page 71: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

iso-volumetric processes of slow rate dissolution of the

mesh serpentine and immediate local precipitation of silica.

They also argue that precipitation of silica is favored at

near-neutral pH conditions. Such conditions differ from

those we assume in our system (high alkali pH). However,

Williams and Crerar (1985) attributed the precipitation of

amorphous silica phases in nature due to the formation of

dense colloids in supersaturate alkaline aqueous solutions.

Fig. 10 Schematic sketch of coupled carbonation–silicification in

dissolution–precipitation processes. a High pH meteoric waters

(yellow arrows) percolating in serpentinite porosity start to dissolve

the serpentine. b In zones of high fluid flow (see the text for more

details), dissolution of the serpentine surface releases Mg2? cations

(white arrows) that migrate to zones of reduced fluid flow, leaving

behind a Si-rich layer. Iron released during this dissolution step

immediately precipitates as magnetite due to local oxidizing condi-

tions favored by the constant renewal of water that characterizes

zones of high fluid flow (Andreani et al. 2009). c Magnesite

precipitates at the expense of serpentine surface in zones of reduced

fluid flow due to local supersaturation and more reducing conditions.

Released silica may precipitate as amorphous silica interstitially to

magnesite aggregates, but increasing pH as the carbonation proceeds

leads to its solubilization and subsequent migration in the serpentine

mesh (orange dashed arrows). d Silica, in aqueous form, propagates

in the serpentine mesh and finally precipitates in amorphous opal-A

subsequently evolving to opal-CT, chalcedony and ultimately quartz.

Complete serpentine replacement in a magnesite ? silica assemblage

may occur. Alternatively, the growing of magnesite may lead to the

closure of initial porosity, preventing additional fluid circulations and

thus ending the carbonation process before going to completion

(Hovelmann et al. 2012)

952 Page 14 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 72: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

In addition, progressive dissolution of silica causes pH of

the solution to drop (Williams and Crerar 1985; Williams

et al. 1985). Following these observations, we infer that

such process may favor the dissolution of the serpentine

mesh and the subsequent grain-by-grain replacement by

amorphous silica phases, as postulated by Lacinska and

Styles (2012). Such a process involves the removal of

substantial amount of magnesium. As shown above, ser-

pentine grain carbonation occurs stoichiometrically, i.e.,

magnesium released after serpentine grain dissolution, then

immediately precipitates as magnesite. Thus, it is likely

that the magnesium released during serpentine mesh silic-

ification migrates out of the reaction zone. This assumption

is consistent with the study of Boschi et al. (2009) that

reported similar magnesium mobilization after serpentine

dissolution and silica precipitation.

Silica evolution

Raman spectroscopy shows that silica formed in response

to serpentine carbonation crystallizes as opal-CT and

chalcedony (Fig. 5). Such silica polymorphs are commonly

described in association with magnesite and are consistent

with formation at low-temperature conditions (Boydell

1921; Bodenlos 1950; Dabitzias 1980; Pohl 1990; Abu-

Jaber and Kimberley 1992b; Klein and Garrido 2011).

According to Lacinska and Styles (2012), the formation of

opal-CT indicates precipitation from supersaturated fluids.

Williams et al. (1985) and Williams and Crerar (1985)

show that the precipitation of silica polymorphs is driven

by multiple steps of dissolution–precipitation. Systemati-

cally, studies made on the silica diagenesis report that

saturated silica solutions do not form opal-CT directly, but

follow a sequential crystallization with first the precipita-

tion of amorphous opal-A (Graetsch et al. 1985; Williams

et al. 1985; Williams and Crerar 1985; Heaney 1993;

Lacinska and Styles 2012). Opal-A then evolves by means

of dissolution–precipitation with concurrent ordering of the

structure and removal of water, forming a pathway as

follow: opal-A (amorphous) ? opal-CT (cristobalite-trid-

ymite assemblage) ? chalcedony ? quartz. According to

Williams et al. (1985), the relationship between solubility

and surface area or particle size is sufficient to explain such

evolution. In our samples, only opal-CT and chalcedony

have been identified so far. One possible explanation to

account for the absence of opal-A is the complete

replacement by opal-CT. The association with magnesite

supports this hypothesis, since carbonates are thought to

enhance the formation of opal-CT (Williams and Crerar

1985). Quartz has never been described associated with

serpentinite in New Caledonia, so it is likely that the

transition chalcedony to quartz does not occur. This reac-

tion is very slow and is more likely to occur in a closed

system, from the precipitation of fluids undersaturated in

silica with respect to opal-A, opal-CT or chalcedony (Lund

1960 and references therein; Williams et al. 1985; Wil-

liams and Crerar 1985). In our samples, vugs containing

chalcedony are systematically characterized by the pre-

sence of a hole on its center (Fig. 4). Similar observations

were reported by Lund (1960) from silicified corals where

both chalcedony and quartz precipitated. This author con-

cluded that the hole served as a conduit for the continuous

circulation of dissolved silica, resulting in the precipitation

of chalcedony. At the opposite, when no hole was

observed, chalcedony was completely replaced by quartz.

As an example, the concurrent reordering that occurs

during the diagenetic pathway of silica is illustrated by the

behavior of chalcedony (Figs. 4 and electronic supplement,

Figure S1). Chalcedony, as many of the microcrystalline

SiO2 varieties, consists of an intimately intergrowth of

a-quartz and moganite. Moganite is a silica polymorph that

typically contains up to 3 wt% of water which is not a

constituent of the structure. Using Raman spectroscopy,

Pop et al. (2004) showed that during the opal-CT to chal-

cedony transition, moganite starts growing after a-quartz

and preferentially in the most crystallized areas. It can be

used as a proxy to evaluate the ordering of the chalcedony

during its transition from opal-CT to quartz (Gotze et al.

1998; Rodgers and Cressey 2001; Pop et al. 2004). In

Fig. 4c, both silica #2 and silica #3 have been identified as

chalcedony but differ in microscopic observation by their

colors. In Raman spectroscopy, silica #2 and silica #3

almost display the same patterns, except the intensity of the

band at 501 cm-1 (moganite) that is higher in the white

chalcedony (electronic supplement, Figure S1). The

moganite content (in wt%) of chalcedony can be calculated

using the Raman band integral ratios I(501)/I(465) (i.e.,

moganite/quartz) and applying the calibration curve pro-

posed by Gotze et al. (1998). We find that both chalcedo-

nies are dominated by moganite, which represents 77 wt%

in brown chalcedony (silica #2, Fig. 4c) and 81 wt% in

white chalcedony. The enrichment of moganite traduces

the progressive ordering of chalcedony which ultimately

transforms in quartz given sufficient time (Williams and

Crerar 1985; Heaney and Post 1992; Rodgers and Cressey

2001; Lynne et al. 2007).

Coupled carbonation–silicification in dissolution–

precipitation processes: a summary

Based on our results and those from previous experi-

mental studies, we infer that serpentine carbonation

occurred due to the circulation of high pH meteoric

waters dissolving the serpentine. Serpentine dissolution

started at grain boundaries and in large grain fractures

that correspond to regions where the fluid flow is the

Contrib Mineral Petrol (2014) 167:952 Page 15 of 19 952

123

Page 73: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

highest (Fig. 10a). In these regions, intensive exchanges

between H? and Mg2? ions led to the development of a

Si-rich layer and the precipitation of magnetite at the

mineral/water interface (Fig. 10b). In contrast, magnesite

precipitation was not favored in these zones of high fluid

flow, which are characterized by strong chemical gradi-

ents and local oxidizing conditions (Andreani et al. 2009).

Released magnesium therefore migrated to regions of

reduced fluid flow, where magnesite nucleated at the

expense of serpentine surfaces by a process of dissolu-

tion–precipitation due to local supersaturation and more

reducing conditions (Fig. 10b–c). Here, the growth of

magnesite was mainly fed by magnesium coming from

the dissolution of adjacent serpentine grain. Potentially,

distal contributions may have occurred (i.e., coming from

the laterization, Barnes et al. 1978), since magnesite

incorporate calcium while this element is not concentrated

in serpentine (Table 1). Silica released during the car-

bonation process may first have precipitated in situ as

amorphous silica (opal-A/CT), but progressive pH

increase during the reaction facilitates the silicon solubi-

lization and subsequent migration away from the fluid–

solid interface (Fig. 10c; Schulze et al. 2004; Andreani

et al. 2009; Hovelmann et al. 2011, 2012). Silica was then

able to propagate in the serpentine mesh, silicifying the

latter by precipitating first as an amorphous gel that

progressively orders its crystalline structure to ultimately

evolve as quartz if given sufficient time (Fig. 10c; Wil-

liams and Crerar 1985; Heaney and Post 1992; Rodgers

and Cressey 2001; Lynne et al. 2007). Magnesium lea-

ched during the serpentine mesh carbonation migrated out

from its original reacting zone and potentially fed larger

magnesite deposits in syn-tectonic fractures, as previously

proposed by Boschi et al. (2009) in the Ligurian ophiolite.

Carbonation process ended when the overall serpentine

was converted into an assemblage of magnesite ? silica

or, alternatively, when magnesite precipitation induced

closure of the initial porosity (Fig. 10d; Hovelmann et al.

2012).

In agreement with recent studies from Andreani et al.

(2009), Boschi et al. (2009) or Hovelmann et al. (2011,

2012), our study demonstrates the importance of active

fracturing in the idea of in situ CO2 sequestration.

Although the volume extend expected during serpentine

carbonation may lead to the system clogging, serpentine

mesh silicification involves substantial removal of mag-

nesium. This magnesium may migrate out of the reaction

zone and subsequently precipitates as massive magnesite

veins along the main structural discontinuities such as

those described in New Caledonia by Quesnel et al. (2013).

As previously underlined by Boschi et al. (2009), such a

process may be considered as an alternative, efficient way

for CO2 sequestration.

Acknowledgments The authors gratefully thank Gilles Montagnac

(ENS Lyon) and Marie-Camille Caumon (GeoRessources Nancy) for

their help during Raman analyses. We also thank Valerie Magnin

(ISTerre Grenoble) for her help during l-XRF analyses and Clement

Marcaillou (Eramet-SLN) and Olivier Vidal (ISTerre Grenoble) for

their contributions on the development of the XRF-mapping software.

This work has been possible thanks to the technical assistance of

Koniambo SA. The financial support from Labex ANR-10-LABX-21-

01 Ressources21 (Strategic metal resources of the 21st century) is

gratefully acknowledged. We thank the editor Jochen Hoefs and the

two anonymous reviewers for their careful revisions that helped to

improve this manuscript.

References

Abu-Jaber NS, Kimberley MM (1992a) Origin of ultramafic-hosted

magnesite on Margarita Island, Venezuela. Mineralium Deposita

27(3):234–241

Abu-Jaber N, Kimberley M (1992b) Origin of ultramafic-hosted vein

magnesite deposits. Ore Geol Rev 7(3):155–191

Alexander G, Mercedes Maroto-Valer M, Gafarova-Aksoy P (2007)

Evaluation of reaction variables in the dissolution of serpentine

for mineral carbonation. Fuel 86(1–2):273–281

Andreani M, Luquot L, Gouze P, Godard M, Hoise E, Gibert B (2009)

Experimental study of carbon sequestration reactions controlled

by the percolation of CO2-rich brine through peridotites.

Environ Sci Technol 43(4):1226–1231

Audet MA (2008) Caracterisation petrographique et structurale des

materiaux ophiolitiques du massif de Koniambo (Nouvelle

Caledonie) et des mineralisations nickeliferes supergenes, Un-

pubished Thesis, Universite du Quebec a Montreal et Universite

de la Nouvelle-Caledonie, pp 355

Auzende A, Daniel I, Reynard B, Lemaire C, Guyot F (2004) High-

pressure behaviour of serpentine minerals: a Raman spectro-

scopic study. Phys Chem Miner 31(5):269–277

Bain GW (1924) Types of magnesite deposits and their origin. Econ

Geol 19(5):412–433

Balucan RD, Dlugogorski BZ (2013) Thermal activation of antigorite

for mineralization of CO2. Environ Sci Technol 47:182–190

Barnes I, O’Neil J, Trescases J (1978) Present day serpentinization in

New Caledonia, Oman and Yugoslavia. Geochim Cosmochim

Acta 42:144–145

Bearat H, McKelvy MJ, Chizmeshya AV, Gormley D, Nunez R,

Carpenter RW, Squires K, Wolf GH (2006) Carbon sequestration

via aqueous olivine mineral carbonation: role of passivating

layer formation. Environ Sci Technol 40(15):4802–4808

Beinlich A, Austrheim H (2012) In situ sequestration of atmospheric

CO2 at low temperature and surface cracking of serpentinized

peridotite in mine shafts. Chem Geol 332–333:32–44

Beinlich A, Plumper O, Hovelmann J, Austrheim H, Jamtveit B

(2012) Massive serpentinite carbonation at Linnajavri, N-Nor-

way. Terra Nova 24:446–455

Birsoy R (2002) Formation of sepiolite-palygorskite and related

minerals from solution. Clays Clay Miner 50(6):736–745

Bobicki ER, Liu Q, Xu Z, Zeng H (2012) Carbon capture and storage

using alkaline industrial wastes. Prog Energy Combust Sci

38:302–320

Bodenlos A (1950) Geology of the red mountain magnesite district,

Santa Clara and Stanislaus Counties, California, California.

J Mines Geol 46:223–278

Boschetti T, Toscani L (2008) Springs and streams of the Taro–Ceno

Valleys (Northern Apennine, Italy): reaction path modeling of

waters interacting with serpentinized ultramafic rocks. Chem

Geol 257(1–2):76–91

952 Page 16 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 74: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Boschi C, Dini A, Dallai L, Ruggieri G, Gianelli G (2009) Enhanced

CO2-mineral sequestration by cyclic hydraulic fracturing and Si-

rich fluid infiltration into serpentinites at Malentrata (Tuscany,

Italy). Chem Geol 265(1–2):209–226. doi:10.1016/j.chemgeo.

2009.03.016

Boydell HC (1921) The magnesite deposits of Euboea, Greece. Econ

Geol 16(8):507–523

Cipolli F, Gambardella B, Marini L, Ottonello G, Vetuschi Zuccolini

M (2004) Geochemistry of high-pH waters from serpentinites of

the Gruppo di Voltri (Genova, Italy) and reaction path modeling

of CO2 sequestration in serpentinite aquifers. Appl Geochem

19(5):787–802

Cluzel D, Meffre S (2002) L’unite de la Boghen (Nouvelle-Caledonie,

Pacifique sud-ouest): un complexe d’accretion jurassique. Don-

nees radiochronologiques preliminaires U-Pb sur les zircons

detritiques. Comptes Rendus Geosci 334:867–874

Cluzel D, Aitchison J, Picard C (2001) Tectonic accretion and

underplating of mafic terranes in the Late Eocene intraoceanic

fore-arc of New Caledonia (Southwest Pacific): geodynamic

implications. Tectonophysics 340(1–2):23–59

Cluzel D, Maurizot P, Collot J (2012) An outline of the geology of

New Caledonia; from Permian-Mesozoic Southeast Gondwana-

land active margin to Cenozoic obduction and supergene

evolution. Episodes 35(1):72–86

Crawford A, Meffre S, Symonds P (2003) 120 to 0 Ma tectonic

evolution of the southwest Pacific and analogous geological

evolution of the 600–220 Ma Tasman Fold Belt System. Geol

Soc Aust Spec Publ 22:377–397

Dabitzias S (1980) Petrology and genesis of the Vavdos cryptocrys-

talline magnesite deposits, Chalkidiki Peninsula, northern

Greece. Econ Geol 75(8):1138–1151

Daval D, Sissmann O, Menguy N, Saldi GD, Guyot F, Martinez I,

Corvisier J, Garcia B, Machouk I, Knauss KG, Hellmann R

(2011) Influence of amorphous silica layer formation on the

dissolution rate of olivine at 90�C and elevated pCO2. Chem

Geol 284:193–209

Fallick AE, Ilich M, Russell MJ (1991) A stable isotope study of the

magnesite deposits associated with the Alpine-type ultramafic

rocks of Yugoslavia. Econ Geol 86(4):847–861

Foster L, Eggleton RA (2002) The Marlborough nickel laterite

deposits. In: Roach IC (ed) Presented at the Regolith and

Landscapes in Eastern Australia Conference, CRC LEME, vol 1,

pp 33–36

Gerdemann S, Dahlin D, O Connor W, Penner L (2003) Carbon

dioxide sequestration by aqueous mineral carbonation of mag-

nesium silicate minerals. Greenhouse gas technologies, Elsevier

Science, Amsterdam 8

Ghoneim M, Saleem I, Hamdy M (2003) Origin of magnesite veins in

serpentinites from Mount El-Rubshi and Mount El-Maiyit.

Eastern Desert Egypt, Archiwum Mineralogiczne, pp 41–63

Giammar D, Bruant R, Peters C (2005) Forsterite dissolution and

magnesite precipitation at conditions relevant for deep saline

aquifer storage and sequestration of carbon dioxide. Chem Geol

217(3–4):257–276

Glasser E (1904) Rapport a M. le Ministre des colonies sur les

richesses de la Nouvelle-Caledonie, Dunod

Goff F, Lackner KS (1998) Carbon dioxide sequestering using

ultramafic rocks. Environ Geosci 5(3):89–102

Gotze J, Nasdala L, Kleeberg R, Wenzel M (1998) Occurrence and

distribution of ‘‘moganite’’ in agate/chalcedony: a combined

micro-Raman, Rietveld, and cathodoluminescence study. Con-

tribut Mineral Petrol 133(1):96–105

Graetsch H, Florke OW, Miehe G (1985) The nature of water in

chalcedony and opal-C from Brazilian agate geodes. Phys Chem

Miner 12:300–306

Griffis R (1972) Genesis of a magnesite deposit, Deloro Twp.,

Ontario. Econ Geol 67(1):63–71

Guthrie GD, Carey JW, Bergfeld D, Byler D, Chipera S, Ziock HJ,

Lackner KS (2001) Geochemical aspects of the carbonation of

magnesium silicates in an aqueous medium. NETL Conference

on Carbon Sequestration, 14 pp

Halls C, Zhao R (1995) Listvenite and related rocks: perspectives on

terminology and mineralogy with reference to an occurrence at

Cregganbaun, Co., Mayo, Republic of Ireland. Miner Deposita

30(3):303–313

Hansen L, Dipple G, Gordon T, Kellett D (2005) Carbonated

serpentinite (listwanite) at Atlin. British Columbia: A geological

analogue to carbon dioxide sequestration, The Canadian Miner-

alogist 43(1):225–239

Harrison AL, Power IM, Dipple GM (2013) Accelerated Carbonation

of Brucite in Mine Tailings for Carbon Sequestration. Environ

Sci Technol 47:126–134

Haug TA, Munz IA, Kleiv RA (2011) Importance of dissolution and

precipitation kinetics for mineral carbonation. Energy Procedia

4(C):5029–5036

Heaney P (1993) A proposed mechanism for the growth of

chalcedony. Contrib Miner Petrol 115(1):66–74

Heaney P, Post JE (1992) The widespread distribution of a novel

silica polymorph in microcrystalline quartz varieties. Science

255(5043):441–443

Hovelmann J, Austrheim H, Beinlich A, Anne Munz I (2011)

Experimental study of the carbonation of partially serpentinized

and weathered peridotites. Geochim Cosmochim Acta

75(22):6760–6779

Hovelmann J, Austrheim H, Jamtveit B (2012) Microstructure and

porosity evolution during experimental carbonation of a natural

peridotite. Chem Geol 334:254–265

Jedrysek MO, Halas S (1990) The origin of magnesite deposits from

the Polish Foresudetic Block ophiolites: preliminary d13C and

d18O investigations. Terra Nova 2(2):154–159

Jones BF, Galan E (1988) Sepiolite and palygorskite. In: Bailey SW

(ed) Hydrous phyllosilicates (Exclusive of Micas) reviews in

mineralogy, 19. Mineralogical Society of America, Washington,

pp 631–674

Jurkovic I, Palinkas LA, Garasic V, Palinkas SS (2012) Genesis of

vein-stockwork cryptocrystalline magnesite from the Dinaride

ophiolites. Ofioliti 37(1):1–14

Kelemen PB, Matter J (2008) In situ carbonation of peridotite for CO2

storage. Proc Natl Acad Sci 105:17295–17300

Kelemen PB, Matter J, Streit EE, Rudge JF, Curry WB, Blusztajn J

(2011) Rates and mechanisms of mineral carbonation in

peridotite: natural processes and recipes for enhanced, in situ

CO2 capture and storage. Annu Rev Earth Planet Sci 39:545–576

King HE, Plumper O, Geisler T, Putnis A (2011) Experimental

investigations into the silicification of olivine: implications for

the reaction mechanism and acid neutralization. Am Mineral

96:1503–1511

Klein F, Garrido CJ (2011) Thermodynamic constraints on mineral

carbonation of serpentinized peridotite. Lithos 126(3):147–160

Klein EM, Langmuir CH (1987) Global correlations of ocean ridge

basalt chemistry with axial depth and crustal thickness. J Geo-

phys Res Solid Earth 92(B8):8089–8115

Krevor S, Lackner KS (2011) Enhancing serpentine dissolution

kinetics for mineral carbon dioxide sequestration. Int J Green-

house Gas Control 5(4):1073–1080

Krishnamurti D (1956) Raman spectrum of magnesite. Proc Math Sci

43(4):210–212

Lacinska AM, Styles MT (2012) Silicified serpentinite—a residuum

of a Tertiary palaeo-weathering surface in the United Arab

Emirates. Geol Mag 1(1):1–11

Contrib Mineral Petrol (2014) 167:952 Page 17 of 19 952

123

Page 75: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Lackner KS, Wendt CH, Butt DP, Joyce EL Jr, Sharp DH (1995)

Carbon dioxide disposal in carbonate minerals. Energy

20(11):1153–1170

Lemaire C (2000) Application des spectroscopies vibrationnelles a la

detection d’amiante dans les materiaux et a l’etude des

serpentines, Unpublished PhD, Universite de Paris 7, Paris,

pp 157

Lin F-C, Clemency C (1981) The dissolution kinetics of brucite,

antigorite, talc, and phlogopite at room temperature and pressure.

Am Mineral 66(7–8):801–806

Luce RW, Bartlett RW, Parks GA (1972) Dissolution kinetics of

magnesium silicates. Geochim Cosmochim Acta 36(1):35–50

Lugli S, Torres-Ruiz J, Garuti G, Olmedo F (2000) Petrography and

geochemistry of the Eugui magnesite deposit (Western Pyrenees,

Spain): evidence for the development of a peculiar zebra banding

by dolomite replacement. Econ Geol 95(8):1775–1791

Lund EH (1960) Chalcedony and quartz crystals in silicified coral.

Am Mineral 45:1304–1307

Lynne BY, Campbell KA, James BJ, Browne PRL, Moore J (2007)

Tracking crystallinity in siliceous hot-spring deposits. Am J Sci

307(3):612–641

Maurizot P, Lafoy Y, Poupee M (2002) Cartographie des forma-

tions superficielles et des aleas mouvements de terrain en

Nouvelle-Caledonie, Zone du Koniambo, BRGM Public Report,

(1:33000)

Munoz M, Pascarellib S, Aquilantib G, Naryginac O, Kurnosovc A,

Dubrovinskyc L (2008) Hyperspectral l-XANES mapping in the

diamond-anvil cell: analytical procedure applied to the decom-

position of (Mg, Fe)-ringwoodite at the upper/lower mantle

boundary. High Press Res 28:665–673

O’Hanley DS (1996) Serpentinites: records of tectonic and petrolog-

ical history. Oxford University Press, Oxford, p 270

Ohmoto H, Watanabe Y, Kumazawa K (2004) Evidence from

massive siderite beds for a CO2-rich atmosphere before *1.8

billion years ago. Nature 429:395–399

Oskierski HC, Bailey JG, Kennedy EM, Jacobsen G, Ashley PM,

Dlugogorski BZ (2012) Formation of weathering-derived mag-

nesite deposits in the New England Orogen, New South Wales,

Australia: implications from mineralogy, geochemistry and

genesis of the Attunga magnesite deposit. Mineralium Deposita.

doi:10.1007/s00126-012-0440-5

Pefrov VP, Vakanjac B, Joksimovic D, Zekic M, Lapcevic I (1980)

Magnesite deposits of Serbia and their origin. Int Geol Rev

22(5):497–510

Plumper O, Royne A, Magraso A, Jamtveit B (2012) The interface-

scale mechanism of reaction-induced fracturing during serpent-

inization. Geology 40(12):1103–1106

Podwojewski P (1995) The occurrence and interpretation of carbonate

and sulfate minerals in a sequence of Vertisols in New

Caledonia. Geoderma 65(3–4):223–248

Pohl W (1990) Genesis of magnesite deposits—models and trends.

Geol Runds 79:291–299

Pop D, Constantina C, Tatar D, Kiefer W (2004) Raman spectroscopy

on gem-quality microcrystalline and amorphous silica varieties

from Romania, Studia UBB. Geologia 49(1):41–52

Power IM, Wilson SA, Small DP, Dipple GM, Wan W, Southam G

(2011) Microbially mediated mineral carbonation: roles of

phototrophy and heterotrophy. Environ Sci Technol

45:9061–9068

Prigiobbe V, Hanchen M, Costa G, Baciocchi R, Mazzotti M (2009)

Analysis of the effect of temperature, pH, CO2 pressure and

salinity on the olivine dissolution kinetics. Energy Procedia

1(1):4881–4884

Prinzhoffer A (1981) Structure et petrologie d’un cortege ophioli-

tique: le massif du Sud (Nouvelle-Caledonie): unpublished

thesis. E.N.S.M, Paris 306 pp

Quesnel B, Gautier P, Boulvais P, Cathelineau M, Maurizot P, Cluzel

D, Ulrich M, Guillot S, Lesimple S, Couteau C (2013) Syn-

tectonic, meteoric water-derived carbonation of the New Cale-

donia peridotite nappe. Geology 41:1063–1066

Renforth P, Washbourne CL, Taylder J, Manning DAC (2011)

Silicate production and availability for mineral carbonation.

Environ Sci Technol 45:2035–2041

Rodgers KA, Cressey G (2001) The occurrence, detection and

significance of moganite (SiO2) among some silica sinters.

Mineral Mag 65(2):157–167

Rudge JF, Kelemen PB, Spiegelman M (2010) A simple model of

reaction-induced cracking applied to serpentinization and car-

bonation of peridotite. Earth Planet Sci Lett 291(1–4):215–227

Saldi GD, Daval D, Morvan G, Knauss KG (2013) The role of Fe and

redox conditions in olivine carbonation rates: an experimental

study of the rate limiting reactions at 90 and 150�C in open and

closed systems. Geochim Cosmochim Acta 118:157–183

Schellart W, Lister G, Toy V (2006) A Late Cretaceous and Cenozoic

reconstruction of the Southwest Pacific region: tectonics con-

trolled by subduction and slab rollback processes. Earth Sci Rev

76(3–4):191–233

Schulze R, Hill M, Field R, Papin P, Hanrahan R, Byler D (2004)

Characterization of carbonated serpentine using XPS and TEM.

Energy Convers Manage 45(20):3169–3179

Sevin B, Ricordel Prognon C, Quesnel F, Cluzel D, Lesimple S,

Maurizot P (2012) First palaeomagnetic dating of ferricrete in

New Caledonia: new insight on the morphogenesis and palae-

oweathering of ‘‘Grande Terre’’. Terra Nova 24:77–85

Sherlock R, Logan M (1995) Silica-carbonate alteration of serpent-

inite; implications for the association of mercury and gold

mineralization in Northern California. Explor Min Geol

4(4):395–409

Sipila J, Teir S, Zevenhoven R (2008) Carbon dioxide sequestration

by mineral carbonation: Literature review update 2005–2007,

Abo Akademi Rep, VT, 1, pp 59

Stanger G (1985) Silicified serpentinite in the Semail nappe of Oman.

Lithos 18:13–22

Teir S, Kuusik R, Fogelholm C, Zevenhoven R (2007) Production of

magnesium carbonates from serpentinite for long-term storage of

CO2. Int J Miner Process 85(1–3):1–15

Teir S, Eloneva S, Fogelholm C-J, Zevenhoven R (2009) Fixation of

carbon dioxide by producing hydromagnesite from serpentinite.

Appl Energy 86(2):214–218

Trescases J (1973) Weathering and geochemical behaviour of the

elements of ultramafic rocks in New Caledonia. Bureau Mineral

Res Geol Geophys Dep Mineral Energy Canberra Bull

141:149–161

Ulrich M (2010) Peridotites et serpentinites du complexe ophiolitique

de la Nouvelle-Caledonie, Unpublished Thesis, Universite de la

Nouvelle-Caledonie et Universite de Grenoble, pp 253

Ulrich M, Picard C, Guillot S, Chauvel C, Cluzel D, Meffre S (2010)

Multiple melting stages and refertilization as indicators for ridge

to subduction formation: the New Caledonia ophiolite. Lithos

115:223–236

van Noort R, Spiers CJ, Drury MR, Kandianis MT (2013) Peridotite

dissolution and carbonation rates at fracture surfaces under

conditions relevant for in situ mineralization of CO2. Geochim

Cosmochim Acta 106 (C):1–24

Whattam SA (2009) Arc-continent collisional orogenesis in the SW

Pacific and the nature, source and correlation of emplaced

ophiolitic nappe components. Lithos 113(1):88–114

Whattam SA, Malpas J, Ali J, Smith IE (2008) New SW Pacific

tectonic model: cyclical intraoceanic magmatic arc construction

and near-coeval emplacement along the Australia-Pacific margin

in the Cenozoic. Geochem Geophys Geosyst 9(3). doi:10.1029/

2007GC001710

952 Page 18 of 19 Contrib Mineral Petrol (2014) 167:952

123

Page 76: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Williams LA, Crerar DA (1985) Silica diagenesis, II, general

mechanisms. J Sediment Res 55(3):312–321

Williams LA, Parks G, Crerar DA (1985) Silica diagenesis, I,

solubility controls. J Sediment Petrol 55(3):301–311

Wilson SA, Dipple GM, Power IM, Thom JM, Anderson RG,

Raudsepp M, Gabites JE, Southam G (2009) Carbon dioxide

fixation within mine wastes of ultramafic-hosted ore deposits:

examples from the Clinton Creek and Cassiar chrysotile

deposits, Canada. Econ Geol 104(1):95–112

Yalicin H, Bozkaya O (2004) Ultramafic-rock-hosted vein sepiolite

occurrences in the Ankara Ophiolitic Melange, Central Anatolia,

Turkey. Clays Clay Mineral 52(2):227–239

Zedef V, Russell M, Fallick A, Hall A (2000) Genesis of vein

stockwork and sedimentary magnesite and hydromagnesite

deposits in the ultramafic terranes of southwestern Turkey: a

stable isotope study. Econ Geol 95(2):429

Contrib Mineral Petrol (2014) 167:952 Page 19 of 19 952

123

Page 77: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Conditions of Magnesite formation

in ophiolites

Quesnel, B., Boulvais, P., Gautier, P., Cathelineau, M., John C. M., Dierick M., Agrinier P., Drouillet M. (2016)

Paired stable isotopes (O, C) and clumped isotope thermometry of magnesite and silica veins

in the New Caledonia Peridotite Nappe

Geoch. Cosmochim. Acta,: 183 : 234-249

- 5 -

Page 78: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Paired stable isotopes (O, C) and clumped isotopethermometry of magnesite and silica veins in the

New Caledonia Peridotite Nappe

Benoıt Quesnel a,⇑, Philippe Boulvais a, Pierre Gautier a, Michel Cathelineau b,Cedric M. John c, Malorie Dierick d, Pierre Agrinier d, Maxime Drouillet e

aGeosciences Rennes – OSUR, Universite Rennes1, UMR 6118 CNRS, 35042 Rennes Cedex, FrancebUniversite de Lorraine, CNRS, CREGU, GeoRessources Lab., 54506 Vandoeuvre-les-Nancy, France

cDepartment of Earth Science and Engineering and Qatar Carbonate and Carbon Storage Research Centre (QCCSRC), Imperial

College London, Prince Consort Road, London SW7 2AZ, UKd Institut de Physique du Globe de Paris, Sorbonne Paris Cite, Univ Paris Diderot, UMR 7154 CNRS, F-75005 Paris, France

eService geologique, Konimbo Nickel SAS, 98883 Voh, New Caledonia, France

Received 11 September 2015; accepted in revised form 21 March 2016; available online 25 March 2016

Abstract

The stable isotope compositions of veins provide information on the conditions of fluid–rock interaction and on the originof fluids and temperatures. In New Caledonia, magnesite and silica veins occur throughout the Peridotite Nappe. In this work,we present stable isotope and clumped isotope data in order to constrain the conditions of fluid circulation and the relation-ship between fluid circulation and nickel ore-forming laterization focusing on the Koniambo Massif. For magnesite veinsoccurring at the base of the nappe, the high d18O values between 27.8‰ and 29.5‰ attest to a low temperature formation.Clumped isotope analyses on magnesite give temperatures between 26 �C and 42 �C that are consistent with amorphous sil-ica–magnesite oxygen isotope equilibrium. The meteoric origin of the fluid is indicated by calculated d18Owater values between�3.4‰ to +1.5‰. Amorphous silica associated with magnesite or occurring in the coarse saprolite level displays a narrowrange of d18O values between 29.7‰ and 35.3‰. For quartz veins occurring at the top of the bedrock and at the saprolitelevel, commonly in association with Ni-talc-like minerals, the d18O values are lower, between 21.8‰ and 29.0‰ and suggestlow-temperature hydrothermal conditions (�40–95 �C). Thermal equilibration of the fluid along the geothermic gradientbefore upward flow through the nappe and/or influence of exothermic reactions of serpentinization could be the source(s)of heat needed to form quartz veins under such conditions.� 2016 Elsevier Ltd. All rights reserved.

Keywords: Peridotite; Weathering; Magnesite; Silica; Stable isotopes; Clumped isotope thermometry; Laterite; Low temperature hydrother-malism

1. INTRODUCTION

Carbonate and silica veins occur in various geologicalcontexts, and are the marker of past geological fluid circu-lation. Understanding the conditions of veins formation isof primary importance to decipher the role of fluids in thetransfer of heat and matter, the transport and deposition

http://dx.doi.org/10.1016/j.gca.2016.03.021

0016-7037/� 2016 Elsevier Ltd. All rights reserved.

⇑ Corresponding author.E-mail addresses: [email protected] (B. Quesnel),

[email protected] (P. Boulvais), [email protected] (P. Gautier), [email protected](M. Cathelineau), [email protected] (C.M. John),[email protected] (M. Dierick), [email protected] (P. Agrinier),[email protected] (M. Drouillet).

www.elsevier.com/locate/gca

Available online at www.sciencedirect.com

ScienceDirect

Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 79: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

of metals, and the mechanical properties of shear zones andfaults. Stable isotopes are one of the historical tools used inthe characterization of fluid–rock interactions (Urey, 1947;McCrea, 1950; Epstein et al., 1953; Craig and Boato, 1955;Emiliani, 1966). Theoretically, stable isotopes give access tothe fluid stable isotopic composition and the temperature ofprecipitation of minerals. However, estimates of tempera-ture and fluid composition are interdependent. That is thereason why stable isotopes studies have to be undertakenwith a strong knowledge of the local geology and in associ-ation with complementary methods of fluid characteriza-tion, like fluid inclusions studies or other geothermometers.

The recent development of the carbonate clumped iso-tope thermometer has rapidly been considered as a verypromising tool in Earth Sciences (Eiler, 2007). This methodis based on the thermodynamic phenomenon of ‘‘clumping”which consists in the preferential bonds formation betweenthe heavy isotopes of carbon and oxygen in carbonate min-erals. Unlike conventional carbonate thermometers, theclumped isotope thermometer allows temperatures of for-mation of carbonate to be estimated independently of theinitial fluid isotopic composition. By extension, it can pro-vide an estimate of the d18O value of the parent fluid sincethe d18O of the carbonate and temperatures are measuredindependently on the same aliquot of sample.

For the last two decades, the study of carbonate and sil-ica veins developed in ultramafic rocks has attracted arenewed interest given the implications of this process forpermanent carbon capture and storage through CO2 miner-

alization (Kelemen and Matter, 2008; Kelemen et al., 2011;Oelkers et al., 2008; Ulrich et al., 2014), and in the under-standing of lateritic nickel ore formation (Butt andCluzel, 2013 and references herein). The New CaledoniaPeridotite Nappe hosts one of the largest lateritic nickelore deposit representing around 30% of the global reservesin nickel, thus making New Caledonia the 7th nickel pro-ducer in the world. In New Caledonia as in many otherplaces worldwide (Caribbean, South America, Balkans,Russia, West Africa, South East Asian, Australia) Ni-laterites result from the intense weathering of ultramaficrocks exposed at the surface under hot and humid climate.Their development requires the dissolution of protolithminerals leading to (i) the export of soluble elements and(ii) the in-situ authigenesis of mineral phases hosting theinsoluble elements. In the case of peridotites in New Cale-donia, Si and Mg are exported whereas the laterites areenriched in iron oxi-hydroxides. Ni, with an intermediatebehavior, is concentrated at the base of the lateritic profilewhere it reaches economic concentrations as Ni-bearinggoethite (Trescases, 1975; Freyssinet et al., 2005). The erra-tic nickeliferous high-grade ore is located along fractures; itis known as garnierite, a mix of Ni-rich serpentine, Ni talc-like minerals and Ni-sepiolite (e.g. Faust, 1966; Brindleyand Hang, 1973; Villanova-de-Benavent et al., 2014;Cathelineau et al., 2015b; Fritsch et al., in press) commonlyassociated with silica.

In New Caledonia (Fig. 1), numerous magnesite(MgCO3), quartz and amorphous silica veins occur at

Fig. 1. Simplified geological map of New Caledonia. Ultramafic rocks are from Maurizot and Vende-Leclerc (2009) and laterites are adaptedfrom Paris (1981).

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 235

Page 80: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

different levels in the peridotite nappe and are thought torepresent by-products of the laterization process (Glasser,1904; Trescases, 1975). A per descensum model of fluid cir-culation was proposed in which Si and Mg are exporteddownward along fractures cutting through the nappe(Trescases, 1975).

Many studies have concentrated on the mineralogicalcharacterization of the Ni bearing phases (Brindley andHang, 1973; Villanova-de-Benavent et al., 2014; Dubletet al., 2015; Cathelineau et al., 2015b; Fritsch et al., inpress), but only few studies focused on the understandingof the physical and chemical conditions of Si, Mg and Nimobility. Based on stable isotope analyses (d18O) of Nitalc-like minerals in garnierite from Morocco (Bou Azzer),Ducloux et al. (1993) showed that they could form between25 �C and 100 �C. Part of this large spread in temperature isdue to uncertainties in the oxygen isotope fractionation fac-tors between Ni talc-like minerals, Ni serpentine-like miner-als and water. These authors consider that this temperaturerange is consistent with a low temperature hydrothermalorigin.

For silica occurrences from New Caledonia, Sabouraudand Trichet (1978) focused on fluid inclusions included inquartz. They identified two generations of quartz, respec-tively crystallized at <30 �C and >120 �C. Unfortunately,these samples were collected in soils and represent trans-ported relicts of dismantled quartz occurrences. It is conse-quently difficult to place them in their genetic context.Recently, Fritsch et al. (in press) invoked conditions oflow-temperature hydrothermalism to explain the successionof Ni-bearing phases filling fractures and ending withquartz. A temperature of about 70 �C is estimated forquartz precipitation by analogy with quartz formed insandstones.

For magnesite formed via carbonation of peridotite var-iously serpentinized, on the basis of stable isotopes analyses(C and O), most authors argue that carbonation occurs atlow temperature and that the fluid is of meteoric originbut rarely propose to link carbonation and laterization pro-cesses (Oman: Kelemen et al., 2011; Streit et al., 2012; WestCalifornia: Barnes et al., 1973; Poland: Jedrysek and Halas,1990; Balkans: Fallick et al., 1991; Jurkovic et al., 2012;Greece: Gartzos, 2004; East Australia: Oskierski et al.,2013; New Caledonia: Quesnel et al., 2013). On the otherhand, some authors argue that few carbonation eventslikely occurred at hydrothermal conditions on the basis ofstable isotopes analyses (Balkans: Fallick et al., 1991; Nor-way: Beinlich et al., 2012) and clumped isotope thermome-try (Oman: Falk and Kelemen, 2015).

In this study, we perform stable isotopes analyses onmagnesite and silica veins and clumped isotope analyseson magnesite veins from New Caledonia, with a specialattention to the Koniambo massif (Figs. 1 and 2) wherethe recent development of a mining site allows the accessto numerous and exceptional outcrops. We show how thetwo isotopic methods are complementary and can help toconstrain the conditions of veins formation. We also ques-tion the per descensum model of fluid circulation proposedto explain the vertical distribution of the different mineral-ization in the peridotite nappe.

2. GEOLOGICAL SETTING

New Caledonia is located in the southwest PacificOcean, 1300 km east of Australia (Fig. 1). Peridotites areabundant on the island as a consequence of the Eoceneobduction of a sliver of oceanic lithosphere (Cluzel et al.,2012). The Peridotite Nappe sub-horizontally overlies thesubstratum which is comprised of several volcano sedimen-tary units (Cluzel et al., 2012). The New Caledonian Peri-dotite Nappe is exposed in the ‘‘Massif du Sud” and as aseries of klippes along the northwestern coast (Fig. 1).

Laterite occurs at the top of the peridotites (Fig. 1). Sev-eral planation surfaces attest to distinct episodes of weath-ering and/or epeirogenic movements during the Cenozoic(Latham, 1986; Chevillotte et al., 2006; Sevin et al., 2012).Capping the laterite profile, a ferricrete composed of indu-rated iron oxides (hematite) mainly occurs as relicts 1 to2 m thick. Below the ferricrete, the limonitic level is mainlycomposed of oxi-hydroxides (goethite) and reaches about20 m in thickness. In this level, the structure of the rock istotally erased. At the base of the profile, the saprolitic levelconsists in an intermediate state of alteration between limo-nite and the underlying peridotite. This level reachesaround 30 m thick. The highly fractured structure of theprotolith is preserved. The saprolite level hosts a diffuseNi mineralization as well as an erratic nickeliferous high-grade ore (Cathelineau et al., 2015a,b; Fritsch et al., inpress), known as garnierite veins.

Garnierite occurs as fracture infillings and consists in anassociation of Ni-rich serpentine and Ni talc-like mineralsoften associated with silica (Trescases, 1975; Cluzel andVigier, 2008; Cathelineau et al., 2015b). Some garnieriteoccurrences formed syn-tectonically (Cluzel and Vigier,2008) and occur occasionally as hydraulic breccia where sil-ica cement hosts clasts of serpentine and Ni talc-like miner-als (Myagkiy et al., 2015). Fractures can also be filled solelyby silica.

Beneath the saprolite, the main part of the nappe con-sists of peridotites in which no silica occurs. Peridotitesare highly fractured. Serpentinization is intensive in thedeformed zones. At the base of the nappe, peridotites arepervasively serpentinized and highly sheared, constitutingthe so-called Serpentine Sole.

The sole hosts numerous magnesite veins (Quesnel et al.,2013) commonly associated with amorphous silica (Ulrichet al., 2014). Elsewhere in New Caledonia, silica also occursas infillings of late normal faults of regional extent, as per-vasive silicification at the base of the nappe, locally knownas the ‘‘Mur de silice” and as veins spatially associated tothe Saint Louis and Koum Oligocene granitoid intrusions(Jacob, 1985; Aye et al., 1986; Paquette and Cluzel, 2007).

These occurrences are not studied here as they do notrelate to the supergene alteration system of the KoniamboMassif. This massif (Fig. 2) corresponds to one of the klip-pen occurring along the northwestern coast (Fig. 1). Aseverywhere on the island, the nappe is capped by a highlydissected and partly reworked lateritic profile (Maurizotet al., 2002). At lower elevations, laterites of the westerly-dipping Kafeate plateau probably belong to a younger pla-nation surface (Latham, 1977; Chevillotte et al., 2006).

236 B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 81: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

3. SAMPLING AND SAMPLES

3.1. Strategy

In the Koniambo massif, a recently opened mining siteallows the sampling of the Peridotite Nappe nearly contin-uously from the Serpentine Sole to the lateritic profile(Quesnel et al., 2016). The Serpentine Sole is well exposedalong cross-sections several hundreds of meters long anda few tens of meters high. The intermediate levels of theNappe can be observed along an access road reaching thesummit of the mining face where mining operations pre-sently expose the laterite and the transition to the fracturedperidotite. The veins have been collected at all levels in thePeridotite Nappe in order to get information on the circu-lation of fluids at the nappe scale (�800 m thick) on the1D vertical direction. On each sampling site (Fig. 2), thenature of the structure hosting veins has been describedwhen possible.

To obtain a first estimate on the representativeness ofthe Koniambo Massif with regards to the whole New Cale-donia Peridotite Nappe, some samples have been collectedat other locations within New Caledonia (Fig. 1): magnesiteveins in the nearby Kopeto Massif (Foue peninsula andKone), quartz veins from the coarse saprolite level fromThio Plateau and Moneo and amorphous silica fromNepoui.

3.2. Magnesite

In the Koniambo massif, magnesite occurs mainly asveins in the Serpentine Sole (Fig. 3a, b). Some magnesite

veins occur along fractures up to a few hundred of metersabove this level (sampling sites 3 and 5, Fig. 2 and Table 1).In many cases, magnesite veins display a cauliflower texturewhich attests of a non-oriented growth. Some occurrencesshow a coarse fibrous texture where fibers of monominer-alic magnesite are orthogonal to vein walls (Fig. 3b, c).At the Serpentine Sole level, magnesite veins occur prefer-entially within and along the margins of meter-thicklow-dipping shear zones that are also marked by the devel-opment of serpentine (Fig. 3a, b). These veins are locallyfolded or offset by subsidiary shear planes (Fig. 3a). Mostof the veins occurring outside the main shear zones havean orientation and coarse fibrous texture suggestive oftension gashes opened during the same shear regime(Fig. 3b, c; see also Quesnel et al., 2013). In soils(Fig. 3d), magnesite appears as nodules that can reach afew tens of cm in width. In general, magnesite samplesare mono-mineralic but locally intimately associated withamorphous silica at the Serpentine Sole level.

3.3. Quartz veins and amorphous silica

Quartz and amorphous silica samples have been exam-ined using optical and electronic microscopies (MEB),and the nature of silica polymorphs has been checked byRaman spectroscopy. In the coarse saprolite level and theunderlying bedrock, quartz veins occur as fractures infill-ings. These infillings correspond to the re-opening ofancient fractures commonly filled by serpentine, and insome cases by nickel-rich (or nickel-bearing) serpentineand/or talc-like minerals (Fig. 4c, d, f). They can occur aslarge branched networks of veins (Fig. 4a), as massive veins

Fig. 2. Geological map of the Koniambo massif adapted from Maurizot et al., 2002. Numbers indicate sampling location and are reported inTable 1 for all samples coming from the Koniambo massif.

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 237

Page 82: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

(Fig. 4c) or as thin films (Fig. 4d). Locally, quartz occurs asbrown micro-crystalline quartz infillings (Fig. 4b, f; ‘‘MicQz” in Table 1), generally preceding the clearer yellow,whitish to clear quartz (Fig. 4b, c, d f; ‘‘Qz” in Table 1).The brown color is due to the presence of micron-size inclu-sions of goethite and hematite. Raman analyses of thesevarious types of quartz show systematically three mainbands around 130 cm�1, 205 cm�1 and 465 cm�1.

Two types of amorphous occurrences were studied: (i)brown opal occurs as very late coating or drapery-like on

blocks of the coarse saprolite level, which constitutes arather rare case, (Fig. 4e; ‘‘brown opal” in Table 1); (ii)translucent opal closely associated with magnesite at theSerpentine Sole, as micro-spherule (Koniambo; Fig. 4g, h;first three rows in Table 1) or as microfracture infillings(Nepoui). Raman analyses show a large band between300 and 500 cm�1 and in some case a band at 460 cm�1.

Similar quartz opal samples were collected in other NewCaledonian occurrences for comparison (Fig. 1 andTable 1).

Fig. 3. Field observations of magnesite occurrences. A–C: Field views of magnesite veins occurring at the serpentine sole. A: field viewillustrating relationship between magnesite veins and shear zone. The thickest vein occurs in the main shear band whereas the thinnest occur inC’-type shear bands. B, C: field view illustrating the fibrous texture of magnesite veins. D: field view of nodular magnesite occurring in a soil.Sample numbers are indicated and are reported in Table 1.

238 B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 83: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Tab

le1

D47,d1

8O

andd1

3Cva

lues

ofmag

nesitesamplesan

dd1

8O

values

ofsilica

samples.‘‘n”indicates

thenumber

ofreplicate

samplesan

alyzed

byclumped

isotopethermometry.T(D

47)uncertainty

isestimated

onthebasisofthe1S

Ean

alytical

uncertainty

inmeasuredD47propag

ated

through

theKluge

etal.(201

5)equation.Thed1

8O

ofthefluid

from

whichmagnesiteform

ediscalculated

usingtheclumped

isotopetemperature

andtheoxygenfractionationfactorbetweenmagnesitean

dH

2O

ofZheng(199

9).

Silica

Mag

nesite

Sam

ple

Localization

Nature

d18O

d18O

d13C

D18O

(Silica-

Magnesite)

D47

Standard

error

T�C

(D47)

nd1

8O

fluid

KoniamboMassif

GBMS-Si-1

Koniambo1:

Sole

Mag

nesitevein

+tran

slucentopal

33.1

28.9

�13.5

4.2

0.67

10.01

635

±6

3�0

.6

BmsSi10

Koniambo1:

Sole

Mag

nesitevein

+tran

slucentopal

32.9

BMSGio

15*

Koniambo1:

Sole

Mag

nesitevein

+tran

slucentopal

29.7

29.1

�13.8

GBMS-2-13

Koniambo1:

Sole

Mag

nesitevein

28.3

�14.3

GBMS-3-13

Koniambo1:

Sole

Mag

nesitevein

28.8

�14.3

GBMS-4-13

Koniambo1:

Sole

Mag

nesitevein

29.0

�12.7

GBMS-5-13

Koniambo1:

Sole

Mag

nesitevein

28.2

�14.3

BMSGio

9*Koniambo1:

Sole

Mag

nesitevein

28.6

�14.3

0.69

50.01

226

±4

3�3

.1BMSGio

3*Koniambo1:

Sole

Mag

nesitevein

28.8

�13.8

0.68

90.01

628

±5

4�2

.4BMSGio

12*

Koniambo1:

Sole

Mag

nesitevein

27.8

�15.2

0.67

50.02

433

±8

3�2

.2GVAV-1-13

Koniambo2:

Sole

Mag

nesitevein

28.7

�13.2

GVAV-2-13

Koniambo2:

Sole

Mag

nesitevein

28.8

�15.3

VAV

Gio

9*Koniambo2:

Sole

Mag

nesitevein

27.8

�14.4

0.69

10.02

028

±7

4�3

.4GMAR-2

Koniambo3:

High.Serp.

Mag

nesitevein

28.2

�14.5

0.69

30.00

627

±2

2�3

.2GMAR-3

Koniambo3:

High.Serp.

Mag

nesitevein

28.8

�12.8

GMAR-5

Koniambo3:

High.Serp.

Mag

nesitevein

29.5

�13.6

GMAR-4

Koniambo3:

High.Serp.

Nodularmag

nesitein

soil

29.5

�9.1

CR-1

Koniambo3:

High.Serp.

Nodularmag

nesitein

soil

28.9

�9.4

KNB3a

Koniambo:Sap

rolite

Qz

25.0

MG1a

Koniambo:Sap

rolite

Qz

25.9

MG3c

Koniambo:Sap

rolite

Qz

28.3

MG5a

Koniambo:Sap

rolite

Qz

24.6

FQZ1

Koniambo7:

Sap

rolite

Qz

25.9

Qzpim

1Koniambo7:

Sap

rolite

Qz

26.0

BAV

Qz1

Koniambo7:

Sap

rolite

Qz

25.9

Sij3

Koniambo9:

Sap

rolite

Qz

29.0

SiCl3

Koniambo9:

Sap

rolite

Qz

28.1

SiClBr1

Koniambo9:

Sap

rolite

Qz

28.5

FQZ2

Koniambo10

:Sap

rolite

Qz

25.4

SiCl1

Koniambo10

:Sap

rolite

Qz

27.5

SiCl2

Koniambo10

:Sap

rolite

Qz

26.5

Sij4

Koniambo11

:Sap

rolite

Qz

28.0

TestPit

Koniambo8:

Sap

rolite

Mic

Qz

25.1

KNB3b

Koniambo:Sap

rolite

Mic

Qz

27.6

Sibr4

Koniambo6:

Sap

rolite

Mic

Qz

23.9

(continued

onnextpage)

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 239

Page 84: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Tab

le1(continued)

Silica

Mag

nesite

Sam

ple

Localization

Nature

d18O

d18O

d13C

D18O

(Silica-

Magnesite)

D47

Standard

error

T�C

(D47)

nd1

8O

fluid

Pit20

7hau

tKoniambo7:

Sap

rolite

Mic

Qz

25.3

KonBas

TestPit

Koniambo8:

Sap

rolite

Mic

Qz

25.1

Sibr6

Koniambo9:

Sap

rolite

Mic

Qz

27.5

Sibr2

Koniambo10

:Sap

rolite

Mic

Qz

27.2

SiR

1Koniambo10

:Sap

rolite

Mic

Qz

24.8

BMSSi13

2Koniambo1:

Sole

Brownopal

33.4

Sibrunelate

Koniambo4:

High.Serp.

Brownopal

32.6

ConvSi1

Koniambo4:

High.Serp.

Brownopal

30.4

New

Caledonia

FOUEGio

Fouepeninsula

Mag

nesitevein

29.1

�12.1

Bellevu

eGio

1Kone

Nodularmagnesitein

soil

29.2

�9.1

DECH

1*Kone

Nodularmagnesitein

soil

29.5

�9.1

0.65

20.00

542

±2

31.5

NEP4

Nepoui

Opal

35.3

MONEO

Moneo

Mic

Qz

26.3

Thio

SP5

Thio

Plateau

Mic

Qz

21.8

‘‘Fau

sse

cargneule”

Thio

Plateau

Qz

24.9

BLV3a

Thio

Plateau

Qz

25.2

3TX1a

Thio

Plateau

Qz

24.5

ThPl3

Thio

Plateau

Qz

25.1

ThPl6

Thio

Plateau

Qz

25.9

ThPl9

Thio

Plateau

Qz

25.8

ThPl10

Thio

Plateau

Qz

25.4

ThPl11

Thio

Plateau

Qz

25.5

240 B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 85: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

4. ANALYTICAL TECHNIQUES

4.1. Oxygen and carbon isotopes in magnesite

The isotopic analysis of 21 magnesite samples has beenperformed at the stable isotope laboratory of the Universityof Rennes 1, France. Samples were finely crushed in aboron carbide mortar, then reacted with anhydrous phos-

phoric acid at 75 �C for 24 h. The experimental fractiona-tion factor between magnesite and CO2 isaCO2�Magnesite ¼ 1:009976 at 75 �C (Das Sharma et al.,2002). CO2 was analyzed on a VG OPTIMA triple collectormass spectrometer. In the absence of magnesite standard,in-lab calcite standard (Prolabo Rennes) samples were ana-lyzed together with the magnesite samples under identicalconditions in order to control the general reliability of the

Fig. 4. A, C, D: field view of quartz veins occurring in the coarse saprolite level of Koniambo massif. B, F: sawed sections through two handspecimen showing complex textural relationships between B: brown micro-crystalline quartz and yellow and white quartz and F: brownmicro-crystalline quartz, translucent quartz and clasts of serpentine (in green). G, E: field view of amorphous silica. G: translucent opalassociated with magnesite at the Serpentine Sole. H: hand sample GBMS-SI-1 where magnesite and amorphous silica are cogenetic. E: brownopal occurring as late coating on blocks of the coarse saprolite. Sample locations are indicated by a white star and are reported in Table 1.(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 241

Page 86: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

protocol. The analytical uncertainty is estimated at ±0.3‰for oxygen and ±0.2‰ for carbon by duplication of severalanalyses. The results are reported using the d notation rel-ative to VSMOW for O and to VPDB for C.

4.2. Magnesite clumped isotopes

The clumped isotope value of a carbonate sample can bededucted by measuring CO2 following acidification, andusing the D47 notation (in ‰) that compares the abundanceof the rare 13C–18O bonds within a sample relative to a cal-culated stochastic distribution (Eiler and Schauble, 2004;Affek and Eiler, 2006):

D47 ¼ R47

2R13 � R18 þ 2R17 � R18 þ R13 � ðR17Þ2"

� R46

2R18 þ 2R13 � R17 þ ðR17Þ2� R45

R13 þ 2R17þ 1

#� 1000

ð1Þ

The D47 value is calculated from the measured ratios (Ri)of masses 45, 46 and 47 to mass 44 and by calculating R13

(13C/12C) and R18 (18O/16O) from R45 and R46 assumingrandom distribution. R17 is calculated from R18 assuminga mass-dependent relationship between 18O and 17O. Thehigher the D47 is, the lower the temperature of formationof the mineral (e.g. Ghosh et al., 2006).

Clumped isotope analyses were performed on 7 samplesof magnesite in the Qatar Stable Isotopes Laboratory atImperial College, London, UK. Magnesite was finelycrushed and samples of �5 mg were digested online in acommon acid bath composed of 105% ortho-phosphoricacid at 90 �C during 1 h. The liberated CO2 gas was contin-uously trapped and purified by passage through a conven-tional vacuum line with multiple cryogenic traps andPorapak-Q trap held at �35 �C (Dennis and Schrag,2010) and was analyzed using a Thermo Finnigan MAT-253 gas source mass spectrometer. The analysis protocolfollowed procedures described by Huntington et al. (2009)and Dennis et al. (2011) comprising measurements thatconsisted of 8 acquisitions with 7 cycles per acquisitionand an integration time of 26 s. Each cycle included a peakcenter, background measurements and an automatic bel-lows pressure adjustment aimed at a 15 V signal at mass44. The sample gas was measured against an Oztech refer-ence gas standard (d13C = �3.63‰ VPDB,d18O = �15.79‰ VPDB). Each sample was measured atleast three times using separate aliquots to improve count-ing statistics. An internal standard (Carrara marble,‘‘ICM”, 0.312‰), a published inter-laboratory standard(‘‘ETH3”, Meckler et al., 2014, 0.634‰ before acid frac-tionation correction) and heated gases were analyzed regu-larly in order to correct for non-linearity (Huntington et al.,2009) and to transfer the measured values in the absolutereference frame (Dennis et al., 2011). Finally, an acid frac-tionation factor of 0.069‰ was added to the D47 values inthe absolute reference frame to account for acid fractiona-tion at 90 �C (Guo et al., 2009; Wacker et al., 2013). Theanalytical uncertainties of the D47 measurements were

added by Gaussian error propagation using standard errorof the mean.

4.3. Oxygen isotopes in quartz veins and amorphous silica

The isotopic analysis of 26 silica samples has been per-formed at the stable isotope laboratory of Institut de Phy-sique du Globe de Paris, France. Samples were finelycrushed and reacted with BrF5 as an oxidizing agent over-night in Ni tubes at 550 �C following the method ofClayton and Mayeda (1963). O2 was directly analyzed ona Thermo-Fisher Delta V mass spectrometer. During thecourse of the analyses, the measurements of 14 NBS28quartz standard allowed to control the general reliabilityof the protocol (mean = 9.67 ± 0.1‰, n = 14).

Thirteen other silica samples have been analyzed at thestable isotope laboratory of Geosciences Rennes (Univer-sity of Rennes 1, France). Samples were and finely crushedand O2 was liberated from minerals through reaction withBrF5 at 670 �C overnight following the method ofClayton and Mayeda (1963). O2 was then converted intoCO2 by reaction with hot graphite. Isotopic ratios weremeasured on a VG SIRA 10 triple collector. NBS 28 quartzstandard and in-house A1113 granite standard were rou-tinely analyzed together with samples; the analytical uncer-tainty is ±0.2‰. The O isotope composition is reportedusing the d notation relative to VSMOW.

5. RESULTS

5.1. Magnesite

The stable isotopes compositions of magnesites (Fig. 5a,b and Table 1) are comparable to those obtained byQuesnel et al. (2013). The oxygen isotope compositions ofmagnesites are homogeneous, with d18O values in the rangeof 27.8–29.5‰. The small variation in d18O values does notcorrelate with any structural or textural characteristics ofmagnesite. The d13C values vary between �15.3‰ and�9.1‰ (Fig. 5b, c and Table 1). The highest d13C valuescorrespond to nodular magnesite from soil. The low valuesbetween �15.3‰ and �12.1‰ correspond to veins from theSerpentine Sole or from highly serpentinized peridotite(sample GMAR-5).

The D47 values and clumped isotope temperatures inmagnesite are reported in Table 1 and illustrated inFig. 5d. The detailed clumped isotope data are presentedin the Electronic Supplement S.1. The clumped isotope tem-peratures presented here are calculated using the D47 cali-bration for calcium carbonate of Kluge et al. (2015),because to date, no D47 calibration exists for magnesite.The temperatures range between 26 and 42 �C (Fig. 5dand Table 1) with an average temperature of 30 ± 5 �Cfor the magnesite veins from the Serpentine Sole and a tem-perature of 42 ± 2 �C for a nodular soil magnesite.

5.2. Quartz veins and amorphous silica

The oxygen isotope compositions of quartz veins andamorphous silica display a large range of d18O values

242 B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 87: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

between 21.8‰ and 35.3‰ (Fig. 5a and Table 1). The val-ues higher than 30‰ correspond to translucent or brownopal. Quartz veins from the coarse saprolite level have atotal range of d18O values between 21.8‰ and 29.0‰, mostvalues being between 24‰ and 28.5‰. There is no system-atic difference between quartz and brown micro-crystallinequartz.

6. DISCUSSION

6.1. Conditions of carbonation

6.1.1. Low temperature geochemistry

The oxygen isotope composition of magnesite samplesare included in the range (27.4‰ 6 d18OMagnesite 6 29.7‰),similar to the results of Quesnel et al. (2013). A first rangeof temperature of formation of magnesite veins between�40 �C and 80 �C was estimated by Quesnel et al. (2013)using the oxygen fractionation factor between magnesiteand water of Chacko and Deines (2008) and a d18O rangeof fluid composition between �1‰ and �7‰ which corre-spond to values of rainwater on isolated island at inter-tropical latitude and low elevation (IAEA database:http://www-naweb.iaea.org/napc/ih/IHS_resources_gnip.html).If other oxygen fractionation factors would have been used

(see details in Electronic Supplement S.2 and S.3), the cal-culated temperature range would be lower, the lowest range(�2 �C and 40 �C) being reached with the fractionationfactor of Aharon (1988). Clumped isotope thermometry isparticularly useful in such situation when (i) several distinctfractionation factors do not converge towards the samerange of temperature of mineral formation and (ii) uncer-tainties on the initial fluid composition is particularly large.To calculate the clumped temperature from D47 values,Falk and Kelemen (2015) used the calibration of Bristowet al. (2011) but in our case, we favor the use of the recentlaboratory calibration for calcium carbonate of Kluge et al.(2015) performed in the same laboratory than our study.Actually, for our samples, the clumped isotope tempera-tures calculated using these two calibrations are similar,taking into account the uncertainties of the measurements.The clumped isotope temperatures obtained using thecalibration of Kluge et al. (2015) average around 30 �C,and would suggest that the magnesite-water fractionationfactors of Aharon (1988; Magnesite 1) and Zheng (1999)are the most consistent with our results.

We use the isotopic equilibrium between amorphous sil-ica and magnesite to test the validity of the calculatedclumped temperature of magnesite formation. We were ableto obtain the d18O of magnesite, d18O of amorphous silica

Fig. 5. The light grey color represents samples from the Koniambo massif, those in dark grey are from elsewhere on the island. All the samplelocations are reported in the Figs. 1 and 2 and in Table 1. The isotopic values are given versus VSMOW (Vienna Standard Mean OceanWater) for d18O and versus VPDB (Vienna PeeDee Belemnite) for d13C. A: histogram of d18O values of magnesite samples (veins and nodules)B: d13C versus d18O diagram of magnesite samples. C: histogram of d18O values of quartz veins amorphous silica. D: diagram showing thetemperature calculated from D47 analyses using the calibration of Kluge et al. (2015) as a function of the d18O values of 7 magnesite samples.

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 243

Page 88: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

and D47 on magnesite measurements for one sample(GBMS-Si-1, Fig. 4g, h and Table 1) where magnesiteand amorphous silica are intimately associated and areinterpreted as cogenetic. At the serpentine sole level wheresample GBMS-Si-1 was collected, pervasive developmentof amorphous silica linked to magnesite formation by ser-pentine dissolution is documented (Ulrich et al., 2014).The isotopic fractionation between silica and magnesite(D18OAmorphous silica-Magnesite) measured on this sample is4.2‰. Actually, there is no published isotopic fractionationfactor between these two species. We therefore combine thevarious oxygen fractionation factors between silica formsand H2O and magnesite and H2O available in the literaturein order to obtain the corresponding temperature (Elec-tronic Supplement S.4). Only three different combinationsprovide realistic temperatures, around 25 �C for two ofthem and around 40 �C for the third. On the same sampleGBMS-Si-1, the D47 measurement allows calculating a tem-perature of 35 ± 6 �C (Table 1), broadly consistent withthese temperatures. It thus appears that both methods agreewith each other, despite the lack of a specific D47 calibrationfor magnesite. More generally, the range of clumped iso-tope temperatures between 26 ± 4 and 42 ± 2 calculatedfor magnesite confirms its low temperature of formationin the Koniambo massif.

6.1.2. Meteoric origin of the fluid

Since clumped isotope thermometry allows estimatingthe temperatures of formation of carbonate independentlyof the knowledge of the fluid composition, we can calculatethe d18O of the fluid from which magnesite formed. The factthat magnesite veins are commonly decimeter thick andabundant on the studied outcrops implies that a largeamount of fluid has been involved in their formation. Thissuggests a high fluid/rock ratio during the fluid circulationthrough the nappe. The observation that the d18O values ofmagnesite are homogeneous (Fig. 5a) pleads for such con-ditions. Consequently, we consider that the calculatedd18O value of the fluid represents its initial compositionand not a composition buffered by the host rock duringits circulation through the nappe. Given the consistencybetween temperature estimate using quartz-magnesite equi-librium and clumped isotopes, we use here the oxygen frac-tionation between magnesite and water from Zheng (1999).The calculated fluids for the 7 magnesite samples have d18Ovalues ranging from -3.4‰ to +1.5‰ (Table.1). The 6 neg-ative d18O values have an average of �2.5‰ with a mini-mum at �3.4‰ and a maximum at �0.6‰. Theycorrespond to magnesite veins occurring at the SerpentineSole. The highest value (d18O = +1.5‰) corresponds to anodular magnesite occurring in a recent soil (DECH1).The oxygen isotope compositions of the fluid are homoge-neous and consistent with the present day composition ofrainwater on isolated island at inter-tropical latitude andlow elevation ranging from �7‰ to �1‰ (AIEA database:http://www-naweb.iaea.org/napc/ih/IHS_resources_gnip.htlm). For the nodular magnesite, the positive oxygen fluidcomposition is interpreted as evaporation at surface condi-tions which leaves residual water enriched in 18O. Thehigher clumped temperature of 42 �C calculated for this

sample seems consistent with the slightly higher tempera-ture needed to precipitate magnesite through evaporationof surface waters.

6.1.3. Fractionation in the carbon isotope system

Whereas the d18O values of magnesite veins are homoge-neous and suggest a single meteoric origin of the fluid fromwhich they formed, the large spread of d13C values impliesseveral sources of carbon and/or specific fractionation effectin the carbon isotope system. The highest d13C values,around �9‰, could result from various sources of carbon.An atmospheric CO2 contribution could, at least in part,explain these high values. A biogenic soil source could alsobe dominant. Indeed, this would be consistent with the factthat magnesite with such d13C values are nodular magne-sites from soil and are probably formed in-situ by evapora-tion as previously explained for sample ‘‘DECH1”. Most ofthe magnesite samples from the Serpentine Sole have lowerd13C values between �15.3‰ and �12.1‰. An organic car-bon contribution from surface soil or from decarboxylationof deep seated organic matter-rich sediments could explainsuch low values.

Independently, several resurgences of hyperalkaline fluidwith pH up to 11 related to present day serpeninization(Barnes et al., 1978; Monin et al., 2014) occur in New Cale-donia (Launay and Fontes, 1985). Among the processesthat could explain low values of d13C, kinetic isotopic frac-tionation of C isotopes is known during hydroxylation ofdissolved CO2 when carbonate forms by diffusive uptakeof dissolved CO2 (Dietzel et al., 1992). Such process hasbeen proposed to explain the extreme depletion of 13C mea-sured in various carbonate occurrences, without importanteffect on the oxygen isotope compositions, as calcite traver-tine deposit in Oman (Clark et al., 1992), sediments cemen-ted by calcite from central Jordan (Fourcade et al., 2007)and also for magnesite deposits in Australia (Oskierskiet al., 2013).

Recently, Hill et al. (2014) and Watkins and Hunt (2015)showed that kinetic fractionation in alkaline to hyperalka-line environment have minor effect on the D47 compositionof carbonate. This can lead to underestimate the clumpedisotope temperature of mineral growth by a few �C. Evenif such effect would have existed during magnesite forma-tion, this thus would not change the conclusion that magne-site formed at low temperature, around 30 �C.

6.2. Conditions of quartz veins and amorphous silica

formation

At the island scale, the d18O values of quartz veins andamorphous silica spread over a range of �15‰. The factthat all quartz veins and amorphous silica occur in the peri-dotite and commonly in deformation zones (at the serpen-tine sole or as fractures or faults infillings) suggest thatthey are probably genetically linked to a unique fluid reser-voir. We argued in Section 6.1.2 that an amorphous silicasample (d18O = 33‰) formed in isotopic equilibrium withmagnesite (sample GBMS-Si-1) from a meteoric waterat low temperature. By inference, we consider that all

244 B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 89: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

amorphous silica occurrences formed from a meteoric fluid.For the quartz veins, their structural position (i.e. at thebase of the laterite profile, in the saprolite) and their asso-ciation with garnierite strongly suggest that the fluid fromwhich they formed also have a meteoric origin. Consideringthe low variability of paleo-latitude of New Caledonia dur-ing the 30 last million years (Paleomap project: http://gcmd.nasa.gov/records/PALEOMAP-Serf.html) and byconsequence the limited variation of the isotopic composi-tion of rainwater during that period, variations in tempera-ture are the most likely factor at the origin of the largespread in d18O values recorded by silica species. Given thatquartz veins are commonly decimeter thick and abundantin the coarse saprolite level it is likely that the systemworked under conditions of large fluid/rock ratios, in whichcase fluids keep their initial isotopic compositions duringinteraction with rocks. By consequence, the d18O of silicamaterials directly reflect the temperature of formation, asthe d18O of the fluid is considered as constant. As no oxy-gen isotope compositions of present-day rainwaters inNew Caledonia are available, we use the mean value ofthe d18O values calculated for the meteoric fluid from whichmagnesite veins formed (mean d18Ometeoric-fluid = �2.5‰,see Table.1) to calculate the temperatures of formation ofamorphous silica and quartz veins. The use of variousoxygen fractionation factors between silica and H2O(Electronic Supplement S.5) provide similar range of tem-perature; only the fractionation factor of Bandriss et al.(1998) was unable to provide a temperature consistent withthe oxygen isotopic equilibrium between quartz and magne-site for sample GBMS-Si-1 (See Section 6.1.1 andElectronic Supplement S.4).

The d18O values of amorphous silica samples rangebetween 29.7‰ and 35.3‰, which correspond to tempera-tures between �15 �C and �40 �C. The quartz veins occur-ring in the coarse saprolite level have d18O values between21.8‰ and 29.0‰, which correspond to temperaturebetween �40 �C and �95 �C, with most samples recordingtemperatures between �40 �C and �80 �C.

To summarize, the main information obtainedfrom the isotope systematics on quartz veins from thesaprolitic level is that they formed under conditions oflow-temperature hydrothermalism rather than at surfacetemperatures.

6.3. Heat sources for quartz veins formation

Magnesite and amorphous silica have ranges of oxygencomposition narrower than that of quartz veins (Fig. 5a, c)which suggests more constant temperatures of formation.The clumped isotope temperatures for magnesite formationare between 26 ± 4 �C and 42 ± 2 �C, which is consistentwith the range of temperature estimated for amorphous sil-ica formation between �15 �C and �40 �C. Even if allamorphous silica are not intimately related to magnesite(typically brown opal silica is not), the fact remains thatboth their similar temperature of formation and their nar-row ranges of d18O values suggest that they are geneticallylinked to the same low temperature fluid. FollowingQuesnel et al. (2013), we can thus propose that magnesite

and amorphous silica occurring at different levels of thenappe record the fluid pathway through the nappe whererapid downward circulation was enhanced by activetectonics.

For quartz veins from the coarse saprolite level the largerange of temperature between �40 �C and �95 �C indicatesvariable conditions and distinctly higher temperatures offormation. The range of temperature estimated for thequartz veins from the coarse saprolite level appears partic-ularly high given their structural position. A first possibilityis to invoke upward movements of hot fluids coming fromdepth. Even if it is difficult to estimate the past thickness ofthe nappe at the time of fluid circulation, the present-daytopographic constraints allow estimating a minimum thick-ness. The mean elevation in the island scale is �640 m asreported by Chevillotte et al. (2006) and the Mont Panie,located on the northern part of the East coast, culminatesat 1628 m. Considering a surface temperature of �25 �C,a past thickness of the nappe between �500 and �2000 mand a paleo-gradient of temperature in the peridotite nappeof 20 �C km�1, temperatures of �35–75 �C may have beenreached at the base of the nappe. Rapid upward migrationof fluids equilibrated at these temperatures might explainthe temperatures recorded by the quartz veins. Occurrencesof hydraulic breccia where quartz cements serpentine andgarnierite clasts in the coarse saprolite level (Myagkiyet al., 2015) are consistent with this idea.

The question of mechanism of precipitation of quartz inthe saprolitic structural level remains. The first possibilityconsists in the precipitation of quartz lead by a change ofpH condition. Indeed quartz solubility is function of pHincreasing rapidly from pH > 9. The pH measured in NewCaledonian resurgences of water at the base of the nappeare basic with pH up to 11 (Monin et al., 2014) whereaspH in coarse saprolite level is around �8 (Trescases,1975). This abrupt decrease of pH at the saprolitic interfacecould explain quartz precipitation. Moreover, an additionalmixing with per descensum meteoric water with pH � 7(Trescases, 1975) and at surface temperature is also possibleand could lead to quartz precipitation.

Nevertheless, unlike the conditions of syn-tectonic mag-nesite formation, no clear evidence of syn-tectonic quartzveins formation is recorded and precludes associating thefluid overpressure to tectonic activity. Consequently, evenif the upward fluid circulation linked to fluid overpressuremay be in agreement with our data, it remains difficult toidentify the cause of such large-scale upward migration offluid.

An alternative hypothesis to explain the �40–95 �C tem-perature range of formation of quartz veins at the coarsesaprolite level is to invoke exothermic reactions of serpen-tinization. Indeed, such reactions are known to produceheat (Fyfe, 1974) and may explain elevated temperatureof fluid in ultramafic–hosted hydrothermal systems(Kelley et al., 2001; Lowell and Rona, 2002; Schroederet al., 2002). Even if serpentinization reactions can explainonly a part of the heat contribution for high temperaturehydrothermal systems, they could be sufficient for low tem-perature hydrothermal system like Lost City type (Mevel,2003). In New Caledonia, several present-day resurgences

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 245

Page 90: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

of low-temperature (reaching 37 �C) and hyperalkaline fluidwith pH up to 11 are interpreted as evidence of present-dayserpentinization (Barnes et al., 1978; Monin et al., 2014). Arecent study (Guillou-Frottier et al., 2015) proposes thatsuch reactions could occur in the coarse saprolite duringthe weathering of the peridotite. These authors argue thatserpentinization reactions should trigger convective fluidcirculation by (i) increasing of temperature by several tensof degrees and (ii) increasing of permeability and porosityby fracturation induced by volume expansion (Kelemenand Hirth, 2012). This model could explain the relativelyhigh temperatures identified for the quartz veins formationand the local occurrences of hydraulic breccia where quartzcements serpentine and garnierite (Myagkiy et al., 2015).Our observation and isotopic data are consistent with thismodel.

7. CONCLUSION

On the basis of stable oxygen isotope data and clumpedisotope analyses, we argue that carbonation at the base ofthe nappe and silicification at the saprolite level of theNew Caledonia Peridotite Nappe occurred under differenttemperature conditions. First, stable isotopes and clumpedisotope analyses of magnesite and opal display a narrowrange of d18O values attesting for formation under homoge-neous conditions, at temperature of �25 to 40 �C for mag-nesite and between �15 �C and �40 �C for amorphoussilica. The oxygen isotope compositions estimated for thefluid at the origin of magnesite and amorphous silica areconsistent with meteoric water compositions. Second, thed18O values of quartz veins from the coarse saprolite leveldisplay a large range, which is indicative of formationbetween �40 �C and �95 �C, typical temperatures of lowtemperature hydrothermalism.

These results lead to question the classical per descensummodel of Ni-laterites ore deposit where silica, magnesiumand nickel leached during supergene alteration of peridotiteare exported by downward fluid circulation along fracturescutting through the nappe. Indeed, our results tends toshow that carbonation of the Serpentine Sole and quartzveins formation at the coarse saprolite level are the resultsof two distinct fluid circulation systems. If the low-temperatures conditions of magnesite formation, specifiedhere using clumped isotope thermometry, are consistentwith the tectonically enhanced downward meteoric watercirculation model previously proposed in Quesnel et al.(2013), higher temperatures of quartz veins formationwould imply that a refinement of the classical per descensummodel is needed. Thermal equilibration of the fluid along anormal geothermic gradient through the nappe or influenceof exothermic reactions of serpentinization could explainsuch temperatures. Regardless of the exact explanation, itmeans that nickel lateritic mineralization, with which thequartz veins studied here are associated, developed underconditions of low-temperature hydrothermalism. Thisinference may be of importance with regards to the proper-ties of migration of nickel during mineralization.

ACKNOWLEDGEMENTS

This study constitutes a part of B. Quesnel PhD which is sup-ported financially and technically by Koniambo S.A. This studyhas benefited from equipment purchased by the Qatar Carbonateand Carbonate Storage Research Center sponsored by Qatar Petro-leum, Shell, and the Qatar Science and Technology Park. PB andMC thank Emmanuel Fritsch for the geological tour he organizedin the CNRT framework. Mineralogical investigations carried outat Nancy were done within the framework of Labex Ressources 21(supported by the French National Research Agency through theNational Program ‘‘Investissements d’avenir”, reference ANR-10-LABX-21-LABEX RESSOURCES 21).

APPENDIX A. SUPPLEMENTARY DATA

Supplementary data associated with this article can befound, in the online version, at http://dx.doi.org/10.1016/j.gca.2016.03.021.

REFERENCES

Affek H. P. and Eiler J. M. (2006) Abundance of mass 47 CO2 inurban air, car exhaust, and human breath. Geochim. Cos-

mochim. Acta 70, 1–12.Aharon P. (1988) A stable-isotope study of magnesites from the

rum jungle uranium field, Australia: implications for the originof strata-bound massive magnesites. Chem. Geol. 69, 127–145.

Aye, F., Bonnemaison, M., Jacob, M. (1986) Etude gıtologique etgeochimique des indices et anomalies en Sb, W, As, Au de lacote Est de la Nouvelle-Caledonie-programme CORDET-Rapport final du projet A4 (1983–84). Bureau des RecherchesGeologiques et Minieres report 85 NCL 079 GMX.

Bandriss M. E., O’Neil J. R., Edlund M. B. and Stoermer E. F.(1998) Oxygen isotope fractionation between diatomaceoussilica and water. Geochim. Cosmochim. Acta 62, 1119–1125.

Barnes I., O’Neil J. R. and Trescases J. J. (1978) Present dayserpentinization in New Caledonia, Oman and Yugoslavia.Geochim. Cosmochim. Acta 42, 144–145.

Barnes I., O’Neil J. R., Rapp J. B. and White D. E. (1973) Silica-carbonate alteration of serpentine: wall rock alteration inMercury deposits of the California Coast Ranges. Econ. Geol.68, 388–398.

Beinlich A., Plumper O., Hovelmann J., Austrheim H. andJamtveit B. (2012) Massive serpentinite carbonation at Linna-javri, N-Norway. Terra Nova 24, 446–455.

Bristow T. F., Bonifacie M., Derkowski A., Eiler J. M. andGrotzinger J. P. (2011) A hydrothermal origin for isotopicallyanomalous cap dolostone cements from south China. Nature

474, 68–71. http://dx.doi.org/10.1038/nature10096.Brindley G. W. and Hang P. T. (1973) The nature of garnierites—

I—Structures, chemical compositions and color characteristics.Clays Clay Miner. 21, 27–40.

Butt C. R. and Cluzel D. (2013) Nickel laterite ore deposits:weathered serpentinites. Elements 9, 123–128.

Cathelineau, M., Caumon, M-C., Massei, F., Brie, D., Harlaux, M.(2015a) Raman spectra of Ni–Mg kerolite: effect of Ni–Mgsubstitution on O–H stretching vibrations. Journal of RamanSpectroscopy, Proceedings GEORaman 2014 (11th), St Louis(USA).

Cathelineau M., Quesnel B., Gautier P., Boulvais P., Couteau C.and Drouillet M. (2015b) Nickel dispersion and enrichment atthe bottom of the regolith: formation of pimelite target-like ores

246 B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 91: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

in rock block joints (Koniambo Ni Deposit, New Caledonia).Miner. Deposita, 1–12.

Chacko T. and Deines P. (2008) Theoretical calculation of oxygenisotope fractionation factors in carbonate systems. Geochim.

Cosmochim. Acta 72, 3642–3660. http://dx.doi.org/10.1016/j.gca.2008.06.001.

Chevillotte V., Chardon D., Beauvais A., Maurizot P. and Colin F.(2006) Long-term tropical morphogenesis of New Caledonia(Southwest Pacific): importance of positive epeirogeny andclimate change. Geomorphology 81, 361–375.

Clark I. D., Fontes J. C. and Fritz P. (1992) Stable isotopedisequilibria in travertine from high pH waters: laboratoryinvestigations and field observations from Oman. Geochim.

Cosmochim. Acta 56, 2041–2050.Clayton R. and Mayeda K. (1963) The use of bromine pentaflu-

oride in the extraction of oxygen from oxides and silicates forisotopic analyses. Geochim. Cosmochim. Acta 27, 43–52.

Cluzel D. and Vigier B. (2008) Syntectonic Mobility of SupergeneNickel Ores of New Caledonia (Southwest Pacific). Evidencefrom Garnierite Veins and Faulted Regolith. Resour. Geol. 58,161–170. http://dx.doi.org/10.1111/j.1751-3928.2008.00053.x.

Cluzel D., Maurizot P., Collot J. and Sevin B. (2012) An outline ofthe geology of New Caledonia; from Permian-MesozoicSoutheast Gondwanaland active margin to Cenozoic obductionand supergene evolution. Episodes 35, 72–86.

Craig H. and Boato G. (1955) Isotopes. Annu. Rev. Phys. Chem. 6,403–432.

Das Sharma S., Patil D. J. and Gopalan K. (2002) Temperaturedependence of oxygen isotope fractionation of CO2 frommagnesite–phosphoric acid reaction. Geochim. Cosmochim.

Acta 66, 589–593. http://dx.doi.org/10.1016/S0016-7037 (01)00833-X.

Dennis K. and Schrag D. (2010) Clumped isotope thermometry ofcarbonatites as an indicator of diagenetic alteration. Geochim.

Cosmochim. Acta 74, 4110–4122.Dennis K. J., Affek H. P., Passey B. H., Schrag D. P. and Eiler J.

M. (2011) Defining an absolute reference frame for ‘clumped’isotope studies of CO2. Geochim. Cosmochim. Acta 75, 7117–7131. http://dx.doi.org/10.1016/j.gca.2011.09.025.

Dietzel M., Usdowski E. and Hoefs J. (1992) Chemical and13C/12C- and 12O/16O-isotope evolution of alkaline drainagewaters and the precipitation of calcite. Appl. Geochem. 7, 177–184.

Dublet G., Juillot F., Morin G., Fritsch E., Fandeur D. andBrown, Jr., G. E. (2015) Goethite aging explains Ni depletion inupper units of ultramafic lateritic ores from New Caledonia.Geochim. Cosmochim. Acta 160, 1–15. http://dx.doi.org/10.1016/j.gca.2015.03.015.

Ducloux J., Boukili H., Decarreau A., Petit S., Perruchot A. andPradel P. (1993) Un gıte hydrothermal de garnierites: l’exemplede Bou Azzer, Maroc. Eur. J. Mineral. 5, 1205–1215.

Eiler J. M. (2007) ‘‘Clumped-isotope” geochemistry—The study ofnaturally-occurring, multiply-substituted isotopologues. EarthPlanet. Sci. Lett. 262, 309–327. http://dx.doi.org/10.1016/j.epsl.2007.08.020.

Eiler J. M. and Schauble E. (2004) 18O13C16O in Earth’satmosphere. Geochim. Cosmochim. Acta 68(23), 4767–4777.http://dx.doi.org/10.1016/j.gca.2004.05.035.

Emiliani C. (1966) Isotopic paleotemperatures. Science 154, 851–857.

Epstein S., Buchsbaum R., Lowenstam H. A. and Urey H. C.(1953) Revised carbonate-water isotopic temperature scale.Bull. Geol. Soc. Am. 64, 1315–1326.

Fallick A. E., Ilich M. and Russell M. J. (1991) A stable isotopestudy of the magnesite deposits associated with the Alpine-typeultramafic rocks of Yugoslavia. Econ. Geol. Bull. Soc. Econ.

Geol. 86, 847–861. http://dx.doi.org/10.2113/gsecongeo.86.4.847.

Falk E. S. and Kelemen P. B. (2015) Geochemistry and petrologyof listvenite in the Samail ophiolite, Sultanate of Oman:complete carbonation of peridotite during ophiolite emplace-ment. Geochim. Cosmochim. Acta 160, 70–90. http://dx.doi.org/10.1016/j.gca.2015.03.014.

Faust G. T. (1966) The hydrous nickel-magnesium silicates—thegarnierite group. Am. Mineral. 51, 279–298.

Fourcade S., Trotignon L., Boulvais P., Techer I., Elie M.,Vandamme D., Salameh E. and Khoury H. (2007) Cementationof kerogene-rich marls by alkaline fluids released duringweathering of thermally metamorphosed marly sediments. PartI: Isotopic (C, O) study of the Khushaym Matruk naturalanalogue (central Jordan). Appl. Geochem. 22, 1293–1310.

Freyssinet P., Butt C. R. M., Morris R. C. and Piantone P. (2005)Ore-forming processes related to lateritic weathering. In Eco-

nomic Geology 100th Anniversary (eds. J. W. Hedenquist, J. F.H. Thomson, R. J. Goldfarb and J Richards). EconomicGeology Publishing Company, New Haven, Connecticut, pp.681–722.

Fritsch E. M., Juillot F. A., Dublet G. A., Fonteneau L., FandeurD., Martin E., Caner L., Auzende A.-L., Grauby O. andBeaufort D. (in press) An alternative model for the formationof hydrous Mg/Ni layer silicates (‘‘deweylite”/‘‘garnierite”) infaulted peridotites of New Caledonia: I. Texture and mineral-ogy of a paragenetic succession of silicate infillings. Eur. J.

Mineral. 28(2). http://dx.doi.org/10.1127/ejm/2015/0027-2503,in press.

Fyfe W. S. (1974) Heats of chemical reactions and submarine heatproduction. Geophys. J. Roy. Astron. Soc. 37, 213–215.

Gartzos E. (2004) Comparative stable isotopes study of themagnesite deposits of Greece. Bull. Geol. Soc. Greece 36, 196–203.

Glasser E. (1904) Rapport a M le Ministre des Colonies sur les

richesses minerales de la Nouvelle-Caledonie. Annales des Mines,Paris, p. 560.

Ghosh P., Adkins J., Affek H., Balta B., Guo W., Schauble E. A.,Schrag D. and Eiler J. M. (2006) 13C–18O bonds in carbonateminerals: a new kind of paleothermometer. Geochim. Cos-

mochim. Acta 70, 1439–1456.Guillou-Frottier, L., Beauvais, A., Bailly, L., Wyns, R., Auge, T.,

Audion, A-S. (2015) Transient hydrothermal corrugationswithin mineralized ultramafic laterites. In: Proceedings of the1”th Biennal SGA (Society for Geology Applied to MineralDeposits) meeting, 24–27 August 2015, Nancy, France, vol. 3,1169–1172.

Guo W. F., Mosenfelder J. L., Goddard W. A. and Eiler J. M.(2009) Isotopic fractionations associated with phosphoric aciddigestion of carbonate minerals: insights from first-principlestheoretical modeling and clumped isotope measurements.Geochim. Cosmochim. Acta 73, 7203–7225.

Hill P. S., Tripati A. K. and Schauble E. A. (2014) Theoreticalconstraints on the effects of pH, salinity, and temperature onclumped isotope signatures of dissolved inorganic carbonspecies and precipitating carbonate minerals. Geochim. Cos-

mochim. Acta 125, 610–652.Huntington K. W., Eiler J. M., Affek H. P., Guo W., Bonifacie M.,

Yeung L. Y., Thiagarajan N., Passey B., Tripati A., Daeron M.and Came R. (2009) Methods and limitations of ‘clumped’ CO2

isotope (D47) analysis by gas-source isotope ratio mass spec-trometry. J. Mass Spectrom. 44, 1318–1329.

Jacob M. (1985) Etude geologique, mineralogique et geochimique des

anomalies en As, Sb, Au et W et indices mineralises lies aux

fractures regionals de la cote sud-est de la Nouvelle-Caledonie Ph.D thesis. University Paul Sabatier, Toulouse, No. 3138, 155p.

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 247

Page 92: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Jedrysek M. O. and Halas S. (1990) The origin of magnesitedeposits from the Polish Foresudetic Block ophiolites: prelim-inary d13C and d18O investigations. Terra Nova 2, 154–159.http://dx.doi.org/10.1111/j.1365-3121.1990.tb00057.x.

Jurkovic I., Palinkas L. A., Garasic V. and Strmic Palinkas S.(2012) Genesis of vein-stockwork cryptocrystalline magnesitefrom the Dinaride ophiolites. Ofioliti 37, 13–26.

Kelemen P. B. and Hirth G. (2012) Reaction-driven crackingduring retrograde metamorphism: olivine hydratation andcarbonation. Earth Planet. Sci. Lett. 345–348, 81–89.

Kelemen P. B. and Matter J. (2008) In situ carbonation ofperidotite for CO2 storage. PNAS 105(45), 17295–17300.

Kelemen P., Matter J., Streit E. E., Rudge J. F., Curry W. B. andBlusztajn J. (2011) Rates and mechanisms of mineral carbon-ation in peridotite: natural processes and recipes for enhanced,in situ CO2 capture and storage. Annu. Rev. Earth Planet. Sci.

39, 545–576. http://dx.doi.org/10.1146/annurev-earth-092010-152509.

Kelley D. S., Karson J. A., Blackman D. K., Fruh-Green G. L.,Butterfield D. A., Lilley M. D., Olson E. J., SchrenkM. O., RoeK. K., Lebon G. T. and Rivissigno P. (2001) AT3-60 shipboardparty, An off-axis hydrothermal vent field near the Mid-Atlantic Ridge at 30�N. Nature 412, 145–149.

Kluge T., John C. M., Jourdan A. L., Davis S. and Crawshaw J.(2015) Laboratory calibration of the calcium carbonateclumped isotope thermometer in the 25–250 �C temperaturerange. Geochim. Cosmochim. Acta 157, 213–227. http://dx.doi.org/10.1016/j.gca.2015.02.028.

Launay J. and Fontes J. C. (1985) Les sources thermales de Prony(Nouvelle-Caledonie) et leurs precipites chimiques. Exemple deformation de brucite primaire. Geol. Fr. 1, 83–100.

Latham, M. (1977) On geomorphology of northern and westernNew Caledonian ultramafic massifs. International symposiumon geodynamics in south-west Pacific, Noumea, Technip Parised, 235–244.

Latham, M. (1986) Alteration et pedogenese sur roches ultra-basiques en Nouvelle-Caledonie: Editions de l’ORSTOM(Office de la Recherche Scientifique et Technique Outre-Mer),Collection Etudes et Theses.

Lowell R. P. and Rona P. A. (2002) Seafloor hydrothermal systemsdriven by the serpentinization of peridotite. Geophys. Res. Lett.29(11).

McCrea J. M. (1950) On the isotopic chemistry of carbonates and apaleotemperature scale. J. Chem. Phys. 18, 849–857.

Maurizot, P., Lafoy, Y., and Poupee, M. (2002) Cartographie desformations superficielles et des aleas mouvements de terrain enNouvelle-Caledonie, Zone du Koniambo: Bureau deRecherches Geologiques et Minieres Public Report RP51624-FR, 45 p.

Maurizot, P., Vende-Leclerc, M. (2009) New Caledonia geologicalmap, scale 1/500000, DIMENC-SGNC, BRGM. Explanatorynote by Maurizot, P. and Collo, J.

Meckler A. N., Ziegler M., Millan M. I., Breitenbach S. F. M. andBernasconi S. M. (2014) Long-term performance of the Kielcarbonate device with a new correction scheme for clumpedisotope measurements. Rapid Commun. Mass Spectrom. 28(15),1705–1715.

Mevel C. (2003) Serpentinization of abyssal peridotites at mid-ocean ridges. C.R. Geosciences 335, 825–852.

Monin C., Chavagnac V., Boulart C., Menez B., Gerard M.,Gerard E., Quemeneur M., Erauso G., Postec A., Guentas-Dombrowski L., Payri C. and Pelletier B. (2014) Thelow temperature hyperalkaline hydrothermal system of theProny bay (New Caledonia). Biogeosciences Dicuss. 11, 6221–6267.

Myagkiy, A., Cathelineau, M., Golfier, F., Truche, L., Quesnel, B.,Drouillet, M. (2015) Defining mechanisms of brecciation in Ni-Silicate bearing fractures (Koniambo, New Caledonia). In:Proceedings of the 1”th Biennal SGA (Society for GeologyApplied to Mineral Deposits) meeting, 24–27 August 2015,Nancy, France, vol. 3, 1181–1194.

Oelkers E. H., Sigurdur R. G. and Matter J. (2008) Mineralcarbonation of CO2. Elements 4, 333–337.

Oskierski H. C., Bailey J. G., Kennedy E. M., Jacobsen G., AshleyP. M. and Dlugogoski B. Z. (2013) Formation of weathering-derived magnesite deposits in the New England Orogen, NewSouth Wales, Australia: implications from mineralogy, geo-chemistry and genesis of the Attunga magnesite deposit. Miner.

Deposita 48, 525–541. http://dx.doi.org/10.1007/s00126-012-0440-5.

Paquette J. L. and Cluzel D. (2007) U-Pb zircon dating of post-obduction volcanic-arc granitoids and a granulite-facies xeno-lith from New Caledonia: inference on Southwest Pacificgeodynamic models. Int. J. Earth Sci. 96, 613–622. http://dx.doi.org/10.1007/s00531-006-0127-1.

Paris, J.P. (1981) Geologie de la Nouvelle-Caledonie: un essai desynthese. Memoire du BRGM (Bureau de Recherches Geolo-giques et Minieres) 113.

Quesnel B., Gautier P., Boulvais P., Cathelineau M., Maurizot P.,Cluzel D., Ulrich M., Guillot S., Lesimple S. and Couteau C.(2013) Syn-tectonic, meteoric water-derived carbonation of theNew Caledonia peridotite nappe. Geology 41, 1063–1066.http://dx.doi.org/10.1130/G34531.1.

Quesnel B., Gautier P., Cathelineau M., Boulvais P., Couteau C.and Drouillet M. (2016) The internal deformation of thePeridotite Nappe of New Caledonia: a structural study ofserpentine-bearing faults and shear zones in the KoniamboMassif. J. Struct. Geol. 85, 51–67.

Sabouraud C. and Trichet J. (1978) Confirmation, par l’etude desinclusions intracristallines, de l’origine superficielle de quartzpyramides. Geoderma 20, 191–199.

Sevin B., Ricordel-Prognon C., Quesnel F., Cluzel D., Lesimple S.and Maurizot P. (2012) First palaeomagnetic dating of ferri-crete in New Caledonia: new insight on the morphogenesis andpalaeoweathering of ‘Grande Terre’. Terra Nova 24, 77–85.http://dx.doi.org/10.1111/j.1365-3121.2011.01041.x.

Schroeder T., John B. and Frost R. B. (2002) Geologic implicationsof seawater circulation through peridotite exposed at slow-spreading mid-ocean ridges. Geology 30, 367–370.

Streit E., Kelemen P. and Eiler J. (2012) Coexisting serpentine andquartz from carbonate-bearing serpentinized peridotite in theSamail Ophiolite, Oman. Contrib. Mineral. Petrol. 164(5), 821–837. http://dx.doi.org/10.1007/s00410-012-0775-z.

Trescases, J.J. (1975) L’evolution geochimique super- gene desroches ultrabasiques en zone tropicale: Formation des gise-ments nickeliferes de Nouvelle-Caledonie. Memoires ORSTOM(Office de la Recherche Scientifique et Technique Outre-Mer)78, 259 p.

Ulrich M., Munoz M., Guillot S., Cathelineau M., Picard C.,Quesnel B., Boulvais P. and Couteau C. (2014) Dissolution-precipitation processes governing the carbonation and silicifi-cation of the serpentine sole of the New Caledonia ophiolite.Contrib. Mineral. Petrol. 167, 952.

Urey H. C. (1947) The thermodynamic properties of isotopicsubstances. J. Chem. Soc., 562–581.

Villanova-de-Benavent C., Proenza J. A., Galı S., Garcıa-Casco A.,Tauler E., Lewis J. F. and Longo F. (2014) Garnierites andgarnierites: textures, mineralogy and geochemistry of gar-nierites in the Falcondo Ni-laterite deposit, Dominican Repub-lic. Ore Geol. Rev. 58, 91–109.

248 B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249

Page 93: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Wacker U., Fiebig J. and Schoene B. R. (2013) Clumped isotopeanalysis of carbonates: comparison of two different aciddigestion techniques. Rapid Commun. Mass Spectrom. 27,1631–1642.

Watkins J. M. and Hunt J. D. (2015) A process-based model fornon-equilibrium clumped isotope effects in carbonates. EarthPlanet. Sci. Lett. 432, 152–165.

Zheng Y. F. (1999) Oxygen isotope fractionation in carbonate andsulfate minerals. Geochem. J. 33, 109–126.

Associate editor: Bjorn Jamtveit

B. Quesnel et al. /Geochimica et Cosmochimica Acta 183 (2016) 234–249 249

Page 94: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 95: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Conditions of Magnesite formation

in ophiolites

Quesnel, B., Boulvais, P., Gautier, P., Cathelineau, M., John C. M., Dierick M., Agrinier P., Drouillet M. (2016)

Formation of silica and magnesite veins in the New Caledonian Peridotite Nappe: geometric and stable isotopes data.

SGA Proceedings

- 6 -

Page 96: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Formation of silica and magnesite veins in the New Caledonian Peridotite Nappe: geometric and stable isotopes data. Benoît Quesnel, Philippe Boulvais, Pierre Gautier Géosciences Rennes, Université de Rennes1 Michel Cathelineau Géoressources, Université de Lorraine, Nancy Cédric M. John Imperial College of London, UK Malorie Dierick, Pierre Agrinier Institut de Physique du Globe de Paris Maxime Drouillet Koniambo Nickel SAS, Nouvelle Calédonie Abstract. The New Caledonian Peridotite Nappe hosts one of the largest nickel ore deposit in the world, in association with laterites developed at the expense of the peridotites. According to a per descensum model of fluid circulation, it is proposed that a genetic link exists between this supergene alteration and the numerous silica and magnesite veins found deeper in the nappe. This work, based on geometrical, mineralogical and stable isotope analyses, provides contstraints on the origin of these veins. Magnesite veins formed at low temperatures from meteoric waters. Silica veins display a large range in δ18O values, which suggests that they formed between surface conditions and low-temperature hydrothermalism. These results question the genetic link of some silica veins with supergene alteration and more generally the per descensum model of Ni-laterite ore formation. Keywords. Peridotite, laterite, magnesite, silica, stable isotopes, clumped isotope thermometry 1 Introduction

Laterites are the result of intense weathering of rocks exposed at the surface. Their development requires dissolution of mineral components of the protolith leading to i) the exportation of soluble elements and ii) the in situ neo-formation of mineral phases hosting the insoluble elements. For peridotites, as in the New Caledonia case, Si and Mg are exported whereas laterites are mainly composed of iron oxi-hydroxides. Ni, with an intermediate behaviour, is concentrated at the base of the lateritic profile where it reaches economic concentrations. Knowledge of the conditions of mobility of Si and Mg during the process should thus improve our understanding of lateritic Ni mineralization.

New Caledonia is located in the southwest Pacific Ocean, 1300 km east of Australia (Fig. 1). Peridotites are abundant on the island as a consequence of the Eocene obduction of a sliver of oceanic lithosphere (Cluzel et al., 2012). The New Caledonian Peridotite Nappe is

essentially exposed in the “Massif du Sud” and as a series of klippes along the northwestern coast (Fig. 1).

Figure 1. Simplified geological map of New Caledonia, modified after Cluzel et al. (2012). The laterites are compiled from Paris (1981). Thick laterites occur at the top of each massif. At present, these laterites represent one of the largest nickel ore deposits in the world and are actively worked by mining companies. Figure 2 shows a typical New Caledonian lateritic profile. At the top of the nappe, the laterite sensu stricto, or limonite level, is mainly composed of iron oxi-hydroxides (goethite) and results from the intensive dissolution of all mineral components of the peridotite. Beneath the limonite horizon, the saprolitic level represents an intermediate state of alteration between limonite and fresh peridotite. It hosts a diffuse Ni mineralization as well as an erratic nickeliferous high-grade ore, known as garnierite (a mix of Ni-rich serpentine and talc-like minerals), which is located along fractures. Silica veins also occur in this level, frequently associated with garnierite. Beneath the saprolite, the main part of the nappe consists of poorly altered peridotite. At the base of

Page 97: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the nappe, peridotites are pervasively serpentinized, forming a highly sheared serpentine sole. The sole hosts numerous magnesite (MgCO3) veins.

Figure 2. Schematic log of the New Caledonian Peridotite Nappe and its interpretation in terms of a per descensum model of fluid circulation, modified after Ulrich et al. (2011). Since the work of Glasser (1904), it is suspected that garnierite as well as the silica and magnesite veins are all genetically linked to the development of the laterites. A per descensum model of fluid circulation has been proposed (Fig. 2), in which the supergene alteration of peridotites by meteoric water has led to leaching and the passive downward export of silica and magnesium through fractures cutting through the nappe (Trescases, 1975). In this study, owing to the opening of a new mining site, we obtained access to exceptional outcrops in the Koniambo Massif (Fig. 1). We focused our attention on silica and magnesite veins as possible by-products of the laterization process. 2 Characteristics of veins 2.1 Magnesite veins

In the serpentine sole, magnesite veins occur preferentially within and along the margins of meter-thick low-dipping shear zones that are also marked by the development of polygonal serpentine (Fig. 3a). These veins are frequently offset by subsidiary shear planes (Fig. 3b,c,d) or folded. Most of the veins occurring outside the main shear zones have an orientation consistent with their interpretation as tension gashes opened during the same shear regime. These observations show that at least some of the magnesite veins were emplaced during the intense deformation recorded by the serpentine sole (Quesnel et al., 2013).

Stable isotope analyses (δ18O & δ13C) were

carried out on 35 samples covering the various types of vein described above. They indicate that the oxygen isotope composition of magnesite is homogeneous, with δ18O values in the range of 27.4‰-29.5‰, while its δ13C

Figure 3. Exemples of magnesite veins showing a close relation to deformation in the serpentine sole of the Koniambo Massif. A,B,C, field views of magnesite veins associated with polygonal serpentine. D, thin-section view of a magnesite nodule crosscut by a shear band.

Page 98: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

varies widely (−16.7‰- −8.5‰). The very positive δ18O values of the magnesite veins suggest a low temperature of formation. The large spread of δ13C values suggests at least two sources of carbon. For the lowest values, around -15‰, an organic carbon contribution, from surface soil, can be proposed. Some kinetic effects on the carbon isotope system associated with the high pH value of fluids (e.g. Fourcade et al., 2007) cannot be ruled out, given that such alkaline fluids are well -known in ultramafics systems (Kelemen et al., 2011). Taken together, the data support a meteoric origin of the fluid from which the magnesite originated (Quesnel et al., 2013). Complementary clumped isotope thermometry analyses were performed on 5 samples of magnesite. With this method (Eiler, 2007), the temperatures of formation of samples are obtained independently of the knowledge of the fluid composition. Preliminary results yield a temperature range of magnesite formation between 24±4ºC and 42±2ºC based on the calibration of Kluge et al (2015). Furthermore, knowing this temperature, the original composition of the fluid can be determined. Using the oxygen fractionation factor between magnesite and water of Chacko and Deines (2008), the calculated oxygen isotope composition of the fluid is between -6‰ and -11.5‰. These values are within or slightly lower than the range of present-day composition of rainwater (-7‰≤δ18O≤-1‰) on isolated islands at inter-tropical latitudes and low elevations (the isotopic composition of present-day rain waters is not known for New Caledonia). Therefore, the low temperature and the calculated δ18O of the fluid support a meteoric water origin for the magnesite veins. 2.2 Silica veins Silica veins mostly occur in the saprolitic levels of the weathering profile (Fig. 4a,b). Different types of silica frequently occur within the same vein (Fig. 4c). Oxygen isotope analyses were carried out on 23 samples covering the various types of silica met in the field. The oxygen isotope compositions vary within 5‰ around a mean value of about 28‰. The highest δ18O values (greater than 30‰) correspond to i) translucent amorphous silica closely associated to magnesite in the serpentine sole, and ii) brownish amorphous silica occurring as a late coating on blocks of the coarse saprolite level. The lower values (lower than 27‰) correspond to silica veins occurring as infillings of a complex network of fractures (Fig. 4a,b). These veins contain various silica types but no amorphous silica. The δ18O variations thus correlate with the mineralogical variability of silica. A difference of about 4‰ is obtained between the δ18O of amorphous silica and magnesite veins. Considering that there is no significant fractionation between amorphous silica and quartz, this value is consistent with the theoretical value of the fractionation between quartz and magnesite at near surface temperatures (cf. Zheng, 1999,). Hence, the oxygen isotope composition of amorphous silica is also consistent with a meteoric origin of the fluid. The lower δ18O values of the non amorphous silica veins likely indicate a higher temperature range of formation. Judging from the large range of δ18O values in these veins, these temperatures were probably variable. This attests for a more complex fluid history than during the formation of magnesite.

Figure 4. Exemples of silica veins in the Koniambo Massif. A and B, field views of silica veins occurring in the coarse saprolite level. C, sawed section through a hand specimen showing the complex textural relationships between at least three different types of silica (brownish, yellowish, and white, the latter occuring as a geodic infilling associated with greenish garnierite).

Page 99: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

3 Conclusion

The δ18O and clumped isotope temperatures of the magnesite and δ18O of silica veins from the Koniambo peridotite Massif, New Caledonia, are consistent with a meteoric origin of the fluid from which they formed. Whereas the magnesite formation is consistent with low-temperature water circulation, the large range of oxygen isotope composition recorded in silica veins implies formation under variable conditions, intermediate between surface temperatures and conditions of low-temperature hydrothermalism. These results question the genetic link of some silica veins with supergene alteration and more generally the per descensum model of Ni-laterite ore formation. References Chacko T, Deines P (2008) Theoretical calculation of oxygen

isotope fractionation factors in carbonate systems. Geochimica et Cosmochimica Acta 72:3642–3660 doi:10.1016/j .gca.2008.06.001.

Cluzel, D, Maurizot, P, Collot, J, Sevin, B, (2012 An outline of the geology of New Caledonia; from Permian-Mesozoic Southeast Gondwanaland active margin to Cenozoic obduction and supergene evolution. Episodes 35:72–86

Eiler J M (2007) “Clumped-isotope” geochemistry—The study of naturally-occurring, multiply-substituted isotopologues. Earth and Planetary Science Letters 262:309–327 doi: 10.1016/j.epsl.2007.08.020.

Fourcade S, Trotignon L, Boulvais P, Techer I, Elie M, Vandamme D, Salameh E, Khoury H (2007) Cementation of kerogene-rich marls by alkaline fluids released during weathering of thermally metamorphosed marly sediments. Part I: Isotopic (C, O) study of the Khushaym Matruk natural analogue (central Jordan). Applied Geochemistry 22:1293-1310

Glasser E (1904) Rapport à M le Ministre des Colo- nies sur les richesses minérales de la Nouvelle- Calédonie. Paris, Annales des Mines, 560 p.

Kelemen, P, Matter, J, Streit, E. E, Rudge, J. F, Curry, W. B, Blusztajn, J (2011) Rates and mechanisms of mineral carbonation in peridotite: natural processes and recipes for enhanced, in situ CO2 capture and storage. Annual Review of Earth and Planetary Sciences 39:545–576. doi:10.1146/annurev-earth-092010-152509

Kluge, T., John, C.M., Jourdan, A.-L., Davis, S., Crawshaw, J., (2015) Laboratory calibration of the calcium carbonate clumped isotope thermometer in the 25-250˚C temperature range, Geochmimica et Cosmochimica Acta (available online), doi:10.1016/j.gca.2015.02.028

Quesnel B, Gautier P, Boulvais P, Cathelineau M, Maurizot P, Cluzel D, Ulrich M, Guillot S, Lesimple S, Couteau C (2013) Syn-tectonic, meteoric water-derived carbonation of the New Caledonia peridotite nappe. Geology 41:1063–1066. doi:10.1130/G34531.1

Trescases, J.J., 1975, L’évolution géochimique super- gène des roches ultrabasiques en zone tropi- cale: Formation des gisements nickélifères de Nouvelle-Calédonie. Mémoires ORSTOM (Of- fice de la Recherche Scientifique et Technique Outre-Mer) 78, 259 p.

Ulrich M, Muñoz M, Guillot S, Chauvel C, Cluzel D, Picard C (2011) Weathering effects on the mineralogical and geochemical composition of the New Caledonia ophiolite. Goldschmidt Conference

Zheng Y.-F. (1999). Oxygen isotope fractionation in carbonate and sulfate minerals. Geochemical Journal, 33, p. 109-126.

Page 100: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 101: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genesis and mineralogy of Ni ores

Cathelineau, M., Myagkiy, A., Quesnel, B., Boiron, M.-C., Gautier, P., Boulvais, P., Ulrich, M., Truche, L.,

Golfier, F., Drouillet, M., (2016a).

Multistage crack seal vein and hydrothermal Ni enrichment in serpentinized ultramafic rocks (Koniambo massif,

New Caledonia).

Mineralium Deposita. doi:10.1007/s00126-016-0695-3

- 7 -

Page 102: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

ARTICLE

Multistage crack seal vein and hydrothermal Ni enrichmentin serpentinized ultramafic rocks (Koniambo massif,New Caledonia)

Michel Cathelineau1&AndreyMyagkiy1 & Benoit Quesnel2 &Marie-Christine Boiron1

&

Pierre Gautier2 & Philippe Boulvais2 & Marc Ulrich3& Laurent Truche1 &

Fabrice Golfier1 & Maxime Drouillet4

Received: 15 February 2016 /Accepted: 17 October 2016# Springer-Verlag Berlin Heidelberg 2016

Abstract Sets of fractures and breccia sealed by Ni-rich sili-cates and quartz occur within saprock of the New Caledonianregolith developed over ultramafic rocks. The crystallizationsequence in fractures is as follows: (1) serpentine stage:lizardite > polygonal serpentine > white lizardite; (2) Ni stage:Ni-Mg kerolite followed by red-brown microcrystallinequartz; and (3) supergene stages. The red-brown microcrys-talline quartz corresponds to the very last stage of the Ni se-quence and is inferred to have precipitated within the 50–95 °C temperature range. It constitutes also the main cementof breccia that has all the typical features of hydraulic fractur-ing. The whole sequence is therefore interpreted as the resultof hydrothermal fluid circulation under medium to low tem-perature and fluctuating fluid pressure. Although frequentlydescribed as the result of a single downward redistribution ofNi and Mg leached in the upper part of the regolith under

ambient temperature, the Ni silicate veins thus appear as theresult of recurrent crack and seal process, corresponding toupward medium temperature fluid convection, hydraulic frac-turing and subsequent fluid mixing, and mineral deposition.

Introduction

The genesis of lateritic Ni deposits developed over ultramaficrocks is generally considered as the result of multi-phase devel-opment through a weathering process where most elements(Mg, Ni, and Si) are leached from the surface and areredeposited at the base of the regolith profile (Trescases 1975;Golightly 2010; Butt and Cluzel 2013; Freyssinet et al. 2005).Recent works on lateritic Ni deposits worldwide have empha-sized the complexity in time and space of the distribution of Ni-

Editorial handling: B. Orberger

* Michel [email protected]

Andrey [email protected]

Benoit [email protected]

Marie-Christine [email protected]

Pierre [email protected]

Philippe [email protected]

Marc [email protected]

Laurent [email protected]

Fabrice [email protected]

Maxime [email protected]

1 Université de Lorraine, CNRS, CREGU, GeoRessources Lab.,54506 Vandoeuvre-lès-Nancy, France

2 Géosciences Rennes, Université Rennes 1, UMR 6118 CNRS,35042 Rennes Cedex, France

3 IPGS-EOST, UMR 7516, Université de Strasbourg,Strasbourg, France

4 Service Géologique, Koniambo Nickel SAS, 98883 Voh, NouvelleCalédonie, France

Miner DepositaDOI 10.1007/s00126-016-0695-3

Page 103: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

rich silicates and their possible role in the distribution of Ni inthe saprock or in the overlaying horizons (yellow limonite)(Wells et al. 2009; Dublet et al. 2012). Thus, nickel is thoughtto be redeposited after its leaching from rock-forming mineralsin the upper laterite horizons, either in the fine-grained lateritewhere goethite is the main Ni-bearing mineral phase, the so-called Blateritic ore,^ or below the laterites, in the rocky saprolitelevel, where it occurs in several mineral phases including goe-thite (Chen et al. 2004; Dublet et al. 2012) and silicates(serpentine, talc-like, and sepiolite) (Wells et al. 2009;Villanova-de-Benavent et al. 2014; Cathelineau et al. 2015a;Fritsch et al. 2016). The evolution of goethite, in particular itsdissolution/recrystallization, is thought to have a crucial role inthe downward Ni redistribution and enrichment (Dublet et al.2015; Wells and Ramanaidou 2015). In the saprock and thebedrock below, Ni-rich silicate phases occur frequently as suc-cessive mineral fillings in fractures.

Although a series of detailed crystal chemistry studies on Ni-serpentine and talc-like phases were performed from the 1970sand 1980s by Brindley and co-authors (Brindley and Hang1973; Brindley and Wan 1975; Brindley et al. 1979) by XRDand then by Manceau and Calas (1985, 1986) using EXAFS,little is known about the genetic conditions of formation of theNi silicate occurring in fractures, such as the temperature, thefluid pressure, and the relationships with tectonics. In NewCaledonia, the fracture sets filled by garnierite wereinvestigated in several pioneering works such as those fromLeguéré (1976) and then Fandeur (2009), at Koniambo, andCluzel and Vigier (2008) at the scale of the island. In theKoniambo deposit, Cathelineau et al. (2015a) discriminatedtwo main types of Ni silicate occurrences in fractures withinthe saprock: (i) type 1 ore which represents the veins with sev-eral mineral fillings including serpentine, talc-like minerals, andquartz and (ii) type 2 ore, called target-like, which consists of anexternal rim of green pimelite around an internal zone composedof white kerolite. All these ore types are probably not synchro-nous, and the long-lived processes may probably explain theconfusion establishing the relative timing of mineral crystalliza-tion. The vein-hosted Ni concentration represents local veryhigh Ni values, which were exploited at the beginning of thetwentieth century but which account only as a few percent of thepresent-day ores. However, they are of prime interest as theyrepresent a preconcentration which is fault controlled and isredistributed within several other ore types: (i) target-like ores,thought to be related to hydration-dehydration cycles whichprovoke successive saturation with respect to talc-like phases,also called kerolite, especially the two end members of the solidsolution (pimelite (Ni100)-kerolite (Ni0)) in joints (Cathelineauet al. 2015a), and (ii) the complex red saprolitic ores, whichcontain relics of all early Ni silicates (vein hosted and target-like) and a fine-grained material dominated by oxyhydroxides.

The genesis of the ore type 1, localized in fractures, is stillpoorly understood. Cluzel and Vigier (2008) provide the first

attempt to link tectonics and mineral formation, which empha-sized that a part of the vein fillings might be considered asBsyntectonic.^ However, the relative timing of mineral crys-tallization in the vein and the physical-chemical conditions oftheir formation remain insufficiently studied to define a con-sistent genetic model.

The objective of this paper is therefore (i) to propose amineral sequence for type 1 veins, especially the spatial andtime relationships between silicates: serpentine polymorphsand talc-like minerals, which are the main Ni-bearing silicatesand quartz, and (ii) to evaluate the conditions of paleo-fluidcirculation, which led to their formation. Exceptional samplespreserved from hydrolysis and oxidation were collected in theopen pits from the Koniambo massif in order to establish acomplete sequence of fracture filling. Detailed in situ mineraldetermination using a combination of complementary analyt-ical techniques (optical and electron microscopies, Ramanspectroscopy, electron microprobe, and laser ablation (LA)-ICP-MS) and a new mineral sequence allow to propose anew genetic model of the Ni silicate vein formation.

Geology of the site

Ni deposits in New Caledonia (Fig. 1) represent 11 % of theworld Ni reserves and 8 % of the nickel production in 2015(USGS 2016). These deposits are developed on ultramaficmassifs, which represent relicts of an obducted peridotitenappe (Fig. 1). The Koniambo massif is one of the klippesoccurring along the northwestern coast. The south-westwardemplacement of the Peridotite Nappe is considered to haveoccurred at ca. 35 Ma (Cluzel et al. 2001, 2012; Paquetteand Cluzel 2007; Maurizot et al. 2009). Outcroppingserpentinized peridotites (Fig. 2) were weathered under trop-ical climate generating a thick lateritic cover (Trescases 1975).The Koniambo massif is characterized by steep slopes andsharp reliefs culminating at around 850 m above sea level.Current enhanced chemical and mechanical erosion, throughthe constant water runoff and hydrolysis of the peridotite min-erals (olivine and pyroxene) predominates, in particular, onthe highest topographic levels (Chardon and Chevillotte2006). Laterite facies occur as relicts, sometimes preservedand perched at the highest altitudes, and in other cases ascollapse or creep sediments in the slopes, following themode l s f rom Chardon and Chev i l l o t t e (2006) .Paleomagnetic ages (around 25 Ma) obtained on lateriticduricrust from Tiebaghi area, north of the Koniambo area,suggest that the main laterite formation has taken place ratherearly, from the late Oligocene to the Miocene, but was thensuccessively dismantled (Chardon and Chevillotte 2006;Sevin et al. 2012).

The Koniambo massif has been exploited intermittentlysince the end of the nineteenth century, and at the end of the

Miner Deposita

Page 104: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

twentieth century by Société minière du Sud Pacifique(SMSP), and is now exploited by SMSP-XSTRATA Nickel(Koniambo Nickel SAS). The deposit hosts around 158.6 mil-lion tons of saprolite that grades 2.47 % Ni at around the 2 %cutoff grade (XStrata 2012). It offers good outcrops, e.g., aseries of open pits, old ones, such as the so-called BCarrièredes anciens^ (Belders^ open pit) or the test pit, and new ones,such as Pit Cagou, and other open pits covering from south-east to north-west a great part of the ore deposit (Fig. 2), in-cluding the south-eastern locality named Trazy. Open pits of-fer a good vertical cross section from bedrock to saprock, withthe possibility to observe and sample veins preserved from thesupergene alteration and oxidation in the bedrock below thesaprock. The most representative fractures, in particular thosefrom the elders and Cagou pits, exhibit either monomineralfillings (Fig. 3) or series of various Mg(Ni)-hydrous silicates(Fig. 4) corresponding to a succession of cracking and sealingevents. These fractures belong to ore type 1 and are differentfrom joints, so-called target like with a rim of pimelite aroundMg-kerolite (Cathelineau et al. 2015a).

Analytical methods

Structural description and measurements of most Ni silicatejoints and fractures were carried out in the open pits along

measurement lines, which are the main walls of the open pits.The wall being oriented at random, there is no bias in themeasurement lines, such as preferential counting of specificdirections. Representative samples were taken in orientedfractures. Minerals were identified on polished thin sectionsusing a multi-analytical characterization. Raman spectroscopywas used to identify the serpentine and kerolite series as wellas SiO2 polymorphs; scanning electron microscopy (SEM)and transmission electron microscopy (TEM) were used toexamine the textures and chronological relationships betweenthe minerals. The in situ chemical analyses were performedusing wavelength-dispersive X-ray spectroscopy (WDSSEM) analysis and quantitative electron microprobe analysis.LA-ICP-MS was used to measure trace elements in quartz.

Microscopy

After careful examination of the samples by optical microsco-py using a transmitted and reflected light microscope, SEManalyses were carried out on polished thin sections with aJEOL SEM, equipped with an energy-dispersive spectrometerusing a Si (Li) semiconductor detector and a HITACHIS-4800 SEM at Service Commun de MicroscopiesElectroniques et de Microanalyses X (SCMEM) (Nancy,France), in order to establish the mineral paragenesis.

Fig. 1 a Simplified geology mapof the New Caledonia modifiedafter Cluzel et al. (2001). The lat-erites are compiled from Paris(1981).

Miner Deposita

Page 105: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Major and trace element quantification

Electron probe microanalyses (EPMA) of kerolite were per-formed using a CAMECA SX100 instrument equipped with aWDS at SCMEM (Nancy, France). Si, Al, Fe, Mn, Mg, Co,Ni, and Na were calibrated using natural and synthetic oxidesalbite (Si, Na), Al2O3 (Al), olivine (Mg), hematite (Fe),MnTiO3 (Mn), Co (Co), and NiO (Ni). The analytical condi-tions were the following: a current of 12 nA, an acceleratingvoltage of 15 kV, and a counting time of 10 s (and 30 or 60 sfor Ni in Ni-poor kerolite). The analyses have a spatial reso-lution of 1 μm. The total Fe is presented as FeO. Structuralformulae were calculated on the basis of an O10(OH)2 base, inthe case of talc-like minerals, and anO5(OH)4 base, in the caseof serpentine. The tetrahedral sheets were assumed to be filledwith Si. The octahedral sheets were composed of Ni and Mg.

Trace element analyses in quartz were carried out at theGeoRessources Laboratory (Vandoeuvre-lès-Nancy, France)using a LA-ICP-MS composed of a 193 nm MicroLas ProArF Excimer coupled with the Agilent 7500c quadrupole ICP-MS. Analytical settings for laser ablation are detailed by Leisenet al. (2012) and Lach et al. (2013). Laser ablation was per-formed with a constant 5-Hz pulse rate, with an ablation craterof 60μm in diameter. The ablatedmaterial is transported using aconstant He flow and mixed with Ar in a cyclone coaxial mixer

prior to entering the ICP torch and being ionized. The ions arethen sampled, accelerated, and focused before being separatedand analyzed in the quadrupole mass spectrometer. The follow-ing isotopes were monitored: 24Mg, 27Al, 53Cr, 55Mn, 57Fe,59Co, 60Ni, 64Cu, and 66Zn. Data reduction was carried outfollowing the standard methods of Longerich et al. (1996) andusing Si content—known from prior EPMA analyses—as aninternal standard. External standard calibration was performedwith the synthetic glass (NIST 610–612, Pearce et al. 1997).

Raman spectroscopy

The nature of the serpentine polymorphs and other silicates wasdetermined by Raman spectroscopy performed on rock sampleand thin and polished section. The Raman spectra were recordedusing a LabRAM HR spectrometer (Horiba Jobin Yvon). Theacquisitionwas performedusing a grating of 1800 lines/mm, a slitwidthof50μm,andaconfocalholeof500μmprovidingaspectralresolution of 0.5 to 1 cm−1. The excitation beamwas provided byan Ar+ laser (Stabilite 2017, Spectra Physics, NewportCorporation) at 514.53 nm, and a power of 200 mW focused onthe sample using a ×50 objective. The spot size was <1 μm.Acquisition time, limited by a weak luminescence, ranged be-tween 2 and 6 s. The number of accumulations was set between10 and 30 to optimize the signal-to-noise ratiowithin a reasonable

Fig. 2 Geological map of the Koniambo massif modified after Maurizot et al. (2002)

Miner Deposita

Page 106: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

acquisitiontime.AllMg-Nihydroussilicateswereanalyzedintworegions, thelowwavenumber(100–1200cm−1)andthehighwavenumber regions (3550–3750 cm−1). The main discriminatingbands are those corresponding to the OH stretching within thehigh-frequency region. In the region of OH stretching vibrations,

the four serpentine polymorphs are distinguished (Lemaire 2000;Auzende et al. 2004), as well as the talc-like minerals (kerolite-pimelite series) (Cathelineau et al. 2015b): (i) serpentines: theOHstretchingbands for lizardite are at ~3690 and~3706cm−1; for thepolygonal serpentine as defined by Baronnet and Devouard

Fig. 3 Representative examplesof field occurrence of vein fillingsat Koniambo: a lizardite (test pit);b lizardite with talc (Cagou pit); cpolygonal serpentine (Trazy),generally green, but here bleached(whitish surface color) due torecent supergene alteration inopen pit; and d fractures by filledkerolite and hematite-goethitemicrocrystalline quartz (lowestpart of the test pit).

Fig. 4 Macroscopic features ofvein fillings and representativeRaman spectra of the differentsilicates: a lizardite vein (test pit);b lizardite with talc, replacedpartially into polygonalserpentine and lizardite II (veincorresponding to that of Fig. 3B,Cagou Pit; and c lizardite veinpartially replaced by polygonalserpentine and lizardite II,recracked, and sealed by kerolite(elder pit (sample KON-207-H)).Colored dots indicate the locationof Raman analyses. Lz lizardite,Tlc talc, Plg. Serp. polygonalserpentine, Ni-Mg krl Ni-Mgkerolite. The blue line corre-sponds to the location of polygo-nal and lizardite II microprobeanalyses from Table 1

Miner Deposita

Page 107: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

(1996), at ~3690 and ~3698 cm−1; for chrysotile at ~3689 and~3701 cm−1; and finally, for antigorite, two main bands are cen-tered at ~3661 and ~3698 cm−1 and (ii) talc: the major OHstretching band is around 3675 cm−1 and (iii) talc-like (kerolite-pimelite series): the shape of the OH stretching vibration bandshows a strong evolution with increasing Ni content (noted from0 to100),witha shift toward lowerwavenumber andanevolutionto a complex band structure displaying several maxima with in-creasing theNicontent (Cathelineauet al. 2015b). In themiddleofthesubstitutionrange(aroundNi50)of thekerolite-pimelite series,the spectrum is a complex band cluster between 3620 and3700 cm−1 with threemajor peaks at 3649, 3670, and 3700 cm−1.

Breccia texture

A few representative examples of breccias were described quan-titativelyusing four parameters: (1) clastmorphology, (2) particlesize distribution, (3) composition ofmatrix, and (4) dilation ratio.The clast morphology is related to the complexity of shape ofindividual clasts, while particle size distribution was applied toall plurality of fragments found in each samples. The shape of thefragments and their size distributionwere described using fractalanalysis techniques. There are several methods to characterize aclast shape by their boundary fractal dimension: box counting,structured walk, dilation, and Euclidean distance mapping(Bérubé and Jébrak 1999). In this study, the boundary fractaldimension of fragment morphology, Dr, was quantified usingthe box-counting method, developed specifically for computer-ized fractal analysis and implemented in the plugin BFracLac^(Karperien 1999–2013) for ImageJ (Rasband 1997–2014).Fractal dimension of particle size distribution,Ds,was calculatedusingacumulativedistributioncurveof thefragmentsizesassum-ing the validity of a power law (Turcotte 1986;Blenkinsop 1991)over a certain range of scale where the curve can be fitted by astraight line.Note that thecalculationof local slopeson thegraph,following Bonnet et al. (2001), leads to the same estimate of thefractal dimension Ds. Processing was carried out on DigitalMicroscope VHX-2000 (© Keyence), whereby the clasts andtheir geometrical sizes were identified. The dilation ratio wascalculated as the volume ratio of matrix and fragments. Thesequantitative analyses were then compared with results of reviewof themain brecciation processes (Jébrak 1997).

Results

Mineralogical and chemical characterization of mineralfillings

Fractures are characterized by a succession of mineral fillingsthat always follows the same order from edge to center: blacklizardite (± talc), polygonal serpentine, white lizardite,kerolite, red-brown microcrystalline quartz, and, finally, clear

quartz (Figs. 3, 4, 5, 6, and 7). The whole sequence corre-sponds thus to a centripetal crystallization from wall rock tothe center of the vein. There is no occurrence of similar suc-cession crosscutting this mineral sequence, which indicatesthat the whole sequence represents a single series of fluidevents. Some veins show breccia textures, the most commonhaving a red microcrystalline quartz cement (e.g., Fig. 5b). Insome instances, the fractures exhibit partial mineral sequence:(i) Fig. 4a, b shows a serpentine filling, which was notreopened during the Ni stage, and (ii) Fig. 4c presents a frac-ture filled by serpentine and affected by microfractures filledby Ni-kerolite, but not by red microcrystalline quartz.Complete successions are, however, frequently found suchas that shown in Fig. 6. The fracture width varies from a fewmillimeters up to a few centimeters, rarely reaching a fewdecimeters. Filled fractures can be observed spread over sev-eral meters both in the horizontal and in the vertical directionsof the fracture plane.

The early filling: the serpentine stage (lizardite, polygonalserpentine)

The lizardite-talc stage Along the edges of the fractures occursblacklizardite,asrecognizedbyRamanspectroscopy,resultingofthe pervasive transformation of the peridotite by localized fluidcirculation in thefracture.Thisdistinctive feature is related toboththecolorof the lizardite itself (Figs.3aand4a)and thepresenceoftinymagnetitecrystals,whichare less than tensofmicrons insize.In the harzburgite, millimetric to centimetric crystals of talc areformed after pyroxene,within thewall rocks replacedby lizarditeon both sides of the fractures, as shown in Figs. 3b and 4b. Theformation of lizardite affects the edges of the fracture on a fewcentimeters when fractures aremetric in length. At the bottom ofthe ophiolite nappe close to the main deformation of the sole,fractures are pluridecametric in length, and the width of lizarditealteration zone can reach more than 0.5 m, and movements onfractures attest to local shearing. In the upper part of themassif ataround 800m of altitude, regular sets of fractures are plurimetricto decametric in length and centimetric in width. These observa-tions are coherent with the structural model of Quesnel et al.(2016b), which discriminates two main distinct structural levelswith two deformation styles, e.g. the sole, highly serpentinized,characterized by pervasive shearing and major faults, and theupper part of themassif, less serpentinized, withmore brittle setsof subparallel fractures.

The polygonal serpentine stage Polygonal serpentine is gen-erally dark to light green in color (Fig. 4b) and replaceslizardite. Polygonal serpentine has the following: (i) a typicalRaman spectra characterized at high frequency by a large peakcomposed of two bands centered at 3690 and 3696–3700 cm−1; (ii) textures under SEM reveal radiating sectorboundaries as described by Baronnet and Devouard (1996);

Miner Deposita

Page 108: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

and (iii) under TEM, a typical texture of short cylinders, whichare empty in their center and when observed perpendicular tothe cylinder axis, displays typical rolled sheet silicate such aschrysotile. The main difference with chrysotile is the veryshort length, with typical length/width ratio being around 2to 4, compared to that of chrysotile which reaches valueshigher than 100.

The white lizardite stage Along the main fractures, clasts oflight green polygonal serpentine are then cemented by a whit-ish lizardite serpentine (Figs. 4b, c, 6, and 7c).

The Ni-Mg kerolite The bluish-greenish Ni-rich minerals(Figs. 3d, 4c, and 5) are talc-like (kerolite) as defined byBrindley and Wan (1975) and Brindley et al. (1979). They arecharacterizedbyadspacingof9.5ÅintheTEMdata.TheRamanspectra are sometimes difficult to obtain on Koniambo kerolitesdue to fluorescence, regardlessof theusedwavelength. In thebestcases, they display a complex series of nine bands in the OH

region in between 3620 and 3700 cm−1 as already described byCathelineau et al. (2015b).

The low content of Ni of serpentine group minerals and thelarge Ni/Mg ratio of kerolites are shown in Fig. 8. Lizardite,polygonal serpentine, and white lizardite are chemically close tothe Mg-serpentine end member (Table 1, Fig. 8). FeO contentvaries around 5.7 ± 0.5 % for lizardite and is slightly lower in thepolygonal and white lizardite (4 to 4.5 % FeO) (Table 1). Nicontent is generally very low in lizardite and polygonal serpentinefrom samples taken away from Ni ores, but it may reach a fewpercent in kerolite-bearing samples. In the latter veins(Fig. 4c), the two serpentines (polygonal and whitelizardite) display homogeneous content of 2 to 3 wt%NiO (Fig. 10). The main Ni silicate, Ni-Mg kerolite(talc-like), is characterized by a strong chemical zona-tion, with a NiO content varying from 10 wt% NiO to27 wt% NiO. It is noted that all kerolites show a sig-nificant stoichiometric deviation from talc, with an ex-cess in octahedral Ni (+Mg) occupancy observed in

Fig. 5 Breccia and fracturefillings. a Fracture with polygonalserpentine (lower part of test pit)filled with Ni-Mg kerolite andlater microcrystalline quartz. bBreccia with a microcrystallinequartz cement and clasts oflizardite and Ni-Mg kerolite. cVeinlet of goethite-hematite mi-crocrystalline quartz (sampleKON-207-H, see Fig. 6). d Detailof c. e Fractal dimension of clastsize distribution, Ds, of breccia

Miner Deposita

Page 109: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Fig. 7 Schematic reconstructionof the succession of mineralfillings in a typical vein in whichcomplete mineral succession isobserved (elder open pit; sampleKON-207-H)

Fig. 6 Succession of lizardite(talc)–polygonal serpentine–Mg-Ni kerolite-hematite microcrys-talline quartz and clear quartzfillings along subparallel fissures(elder pit, KON-207-H), withrepresentative Raman spectra ofthe different silicates. Same ab-breviations as in Fig. 4; Mic Qtzgoethite microcrystalline quartz

Miner Deposita

Page 110: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

correlation with a deficiency in the tetrahedral occupan-cy. Kerolites have a Si/(Si + Mg + Ni) ratio around0.52 ± 0.02 distinct of talc (0.4) and serpentine (0.4–0.43) (Fig. 8).

Quartz stages

Red-brown quartz and breccia The kerolite veins, in almostall samples, are fractured and cemented by microcrystallinequartz characterized by an intense red-brown color due toabundant microparticles of iron oxyhydroxides (Fig. 5d). Inthe Raman spectra presented in Fig. 6, the iron oxyhydroxideshows bands around 240, 300, 390, and 550 cm−1, close tothose of goethite, but rather wide. The large band around400 cm−1 frequently found in brown quartz has more similar-ities with that obtained on δ-FeOOH than the band of well-crystallized goethite (de Faria et al. 1997), indicating that theiron oxyhydroxide is probably poorly crystallized. The mainbands typical of hematite at 404 and 1310 cm−1 were notfound. The XRD pattern of bulk red-brown quartz and thatof the finest grain fraction extracted from the red-brown quartzshow only reflections of quartz and indicate that quartz has ahigh crystallinity at variance with the iron oxyhydroxide,which is either not sufficiently abundant or crystalline to bedetected. The texture of quartz grains is very fine grained, butits high crystallinity is revealed by the well-defined and fineRaman bands at 465 and around 205 cm−1 (Fig. 6).

Table 1 Electron probe microanalyses and structural formulae for the main silicates forming the fracture cements: lizardite, polygonal serpentine,kerolite, and microcrystalline quartz

Sample Lizardite Polygonalserpentine

Ni-Mgkerolite

Microcrystallinequartz

Microcrystallinequartz

KON-low testpit

KON-207-H

KON-207-H Si-BR-4 KON-207-H

n = 3 n = 18 n = 22 n = 19 n = 9m σ m σ m σ m σ m σ

% Na2O 0.01 0.01 0.01 0.01 0.00 0.00 0.03 0.02 0.01 0.01

% MgO 37.68 0.38 36.04 1.07 18.57 3.71 1.12 0.81 0.51 0.27

% Al2O3 0.15 0.09 0.12 0.04 0.18 0.22 0.84 0.85 0.30 0.14

% SiO2 42.07 1.64 43.83 0.61 52.98 2.37 83.11 5.62 90.56 3.34

% K2O 0.01 0.00 0.01 0.01 0.05 0.02 0.02 0.02 0.00 0.01

% CaO 0.02 0.01 0.07 0.19 0.00 0.00 0.05 0.03 0.02 0.02

% Cr2O3 0.00 0.00 0.03 0.06 0.02 0.05 0.15 0.07 0.11 0.06

% MnO 0.11 0.03 0.04 0.04 0.02 0.03 0.05 0.04 0.03 0.04

% FeO 4.55 1.00 5.72 0.56 0.19 0.22 10.31 4.63 6.83 3.38

% CoO 0.00 0.00 0.02 0.03 0.01 0.03 0.02 0.04 0.01 0.02

% NiO 0.29 0.01 1.08 0.23 20.80 5.13 1.65 0.29 0.48 0.63

Total 84.90 86.96 0.75 92.82 1.81 96.33 1.65 98.17 0.74

Si 2.03 2.08 3.86

Al 0.00 0.01 0.02

Fe 0.18 0.23 0.01

Mg 2.72 2.55 2.01

Mn 0.01 0.00 0.00

Ni 0.01 0.04 1.23

Fe + Mg + Ni 2.91 2.82 3.25

Structural formulae were calculated on the basis of an O10(OH)2 base, in the case of talc-like, and an O5(OH)4 base, in the case of serpentine

Fig. 8 Si-Mg + Fe-Ni triangular diagram (electron probe microanalyses)for lizardite, polygonal serpentine, and kerolite from the vein (sampleKON-207-H). Same abbreviations are in Fig. 4

Miner Deposita

Page 111: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

At the micron scale, the electron microprobe analyses of themicrocrystalline iron oxyhydroxides bearing quartz indicatethat the iron content is highly variable at small scale due tothe relative abundance of microsize iron oxyhydroxide particles(Fig. 9). The Fe2O3 content in redmicrocrystalline quartz variesbetween 4 and 16 wt%with a mean at 10.3 wt%, indicating thatit is a mineral assemblage composed of quartz at more than85 % and iron hydroxides as minute inclusions of less than1-μm width, probably with a poor crystallinity (Table 1). Inaddition to Fe, Ni, Al, and Mg have mean values of 1.65 wt%NiO, 0.84 wt% Al2O3, and 1.12 wt%MgO, respectively. In thesample displayed in Fig. 6, Ni has a much lower content around0.48 wt% NiO (Table 1) and the other metals are also lower inconcentration (0.3 wt% Al2O3; 0.5 wt% MgO). LA-ICP-MSanalyses of this sample are given in Table 2. The red-brownquartz veins contain significant and rather homogeneousamounts of metals (Al, Mn, Cr), a feature which is never en-countered in other quartz types, such as the clear quartz veinlets.

Late quartz Sets of subparallel thin veinlets of translucent ormilky quartz are parallel to the earlier structures (Fig. 7). Quartz

displays a good crystallinity as shown by thin Raman bands,petrographic features typical of quartz grains of more than tensofmicrons inwidth, and is, inmost cases, transparentwithout anysolid inclusion. Figure 10 summarizes the mineral succession inveins and highlights the three main stages described previously.

Structural and textural characterization of veins

Fracture distribution and orientations Field observationsshow that fractures filled only by kerolite are rarely observed,as kerolite is generally sealing reopen serpentine fractures (main-ly polygonal serpentine). Two types of fractures have been mea-sured on several sites in the area (Fig. 2). The first one consists offractures filled only by polygonal serpentine, while the secondone consists of fractures where kerolite is present together withpolygonal serpentine at the selvage of the kerolite fillings. Manyof these fractures can be considered as fault planes. Severalpolygonal serpentine fillings form accreted slickenfibers andhave staircase geometry attesting for their syntectonic formation(Quesnel et al. 2016b). For the kerolite fillings, some evidenceof slickensides attests, at least in part, for their syntectonic

Fig. 9 NiO and FeO profilethrough the vein shown in Fig. 6(sample KON-207-H). Sameabbreviations are in Fig. 4

Table 2 LA-ICP-MS analyses ofmetal contents in themicrocrystalline quartz (sampleKON-207-H)

Conc. (ppm) Mg Al Cr Mn Fe Co Ni Cu Zn

KON-207-H-3-1 1869 1829 1285 282 33,613 53 2313 8 54KON-207-H-3-2 1914 1926 1316 301 35,066 56 2415 9 57KON-207-H-3-3 1970 1974 1383 310 36,467 60 2368 8 57KON-207-H-4-1 2153 1677 1178 265 30,938 49 2238 7 47KON-207-H-4-2 1743 2000 1423 320 37,243 61 2404 9 56KON-207-H-4-3 1743 2000 1423 320 37,243 61 2404 9 56Mean 1899 1901 1335 300 35,095 57 2357 8 54Standard deviation 155 127 95 22 2473 5 69 1 4

Location of analyses close to that of the Raman spectrum from Fig. 6

Miner Deposita

Page 112: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

formation (Cluzel and Vigier 2008). To summarize, (i) only fewevidence of slickensides associated to kerolite is observed on thefield and (ii) this type of structure does not provide kinematiccriteria on the fault plane. Consequently, the stereograms pre-sented in Fig. 11 only present the orientations of the fracturesassociated with polygonal serpentine and with kerolite, withoutkinematic interpretation. The two stereograms show a largeoverlap in the distribution of the poles associated to polygonalserpentine and kerolite. Kerolite fractures display predominantdirections in between N170 and N20. For the polygonal frac-tures, the predominant directions are in between N180 and N30,and fractures show a regular spacing. The two populations offractures are steeply dipping between 70° and 90°.

For the kerolite population, two clusters of poles differfrom the overlap area. They are located in the 0°–45° and

180°–225° quarter of the stereogram and correspond, in part,to the fracture filled by kerolite with no polygonal serpentine.

Characterization of breccia Breccia samples (Fig. 5) arecomposed of broken, angular-shaped fragments of (i) serpen-tine (lizardite, partially to totally converted to polygonal ser-pentine) and (ii) bluish green Ni-Mg talc-like, cemented to-gether by a reddish-brown microcrystalline quartz-goethitematrix. The dilation ratio of the samples is high: the matrixis abundant and represents more than 80 % of the breccia.Jigsaw textures are locally recognized at grain scale: frag-ments still match along adjacent surfaces and show that theyare former parts of larger clasts. The matrix of the brecciasamples is characterized by simple and homogeneous fillingmaterial that consists of two co-precipitated minerals: (i) mi-crocrystalline quartz and (ii) microparticles of goethite.

Fragments from the breccias have an angular, nearlyEuclidean shape, with a fractal dimension Dr = 1.034. The par-ticle size distribution (PSD) of 209 clasts is described by a powerlaw relationship. The slope of best-fit line on a log-log plot ofN(>r) vs. r is equivalent to the fractal dimension, Ds, of the PSD(Turcotte 1986) and found to be 1.144 (Fig. 5e). Particles largerthan 2834 μm deviate from the power law relationship becauseof censoring effects (Bonnet et al. 2001). These effects implythat fragments with sizes comparable to the sampled region tendto deviate from the power law distribution.

Discussion

The mineral sequence determined on main Ni silicate fracturesfollows always the same order: (i) serpentine stage: lizardite >polygonal serpentine > white lizardite, (ii) Ni stage: Ni-Mg

Fig. 10 Parageneticsequenceshowing the threemainstagesof fracture filling

Fig. 11 Stereograms (lowerhemisphere, equal-area projec-tion) showing the orientation ofthe main fractures characterizedby a polygonal serpentine afterlizardite and b veins with keroliteinfilling using the Stereonet soft-ware (Allmendinger et al. 2013).Data were collected on seven lo-calities (Fig. 2). Shades of grayvary as a function of the proba-bility for the presence of the polesfollowing the Kamb method

Miner Deposita

Page 113: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

kerolite followed by red-brownmicrocrystalline quartz, and (iii)later stages including the formation of the target-like ores, asjoint fillings, and milky to translucent quartz. The goethite-richmicrocrystalline quartz corresponds to the last event of the Nistage and constitutes also the main cement of breccia. Thismicrocrystalline quartz stage is thus of primary importance forunderstanding the sequence of fracture fillings, as it representsan excellent proxy to infer the conditions of the associated talc-like formation. Several indications can be deduced from themicrocrystalline quartz stage closely associated with the talc-like: (i) the geometric distribution of fractures and conditions ofbreccia formation, (ii) redox conditions, and (iii) temperature ofprecipitation based on stable isotope data on quartz. The fol-lowing paragraphs therefore examine these three main points inorder to provide new insights into the Ni silicate vein formationand by extension into the genetic model of Ni ore formation.

Fracture opening and brecciation

Two types of deformation textures have been found: (i) wide-spread series of fractures where successive fillings are nearlyparallel to the previous one resulting of successive crackingand sealing and (ii) fracture opening with subsequent brecci-ation, found locally.

For the cracked and sealed fractures, the main observationsare the following: kerolite-filled small fractures which post-date those of the polygonal serpentine generally correspond tothe new cracking of a fractured zone containing serpentines.This suggests a strong influence of the inherited fracture net-work on the fluid circulation pathway at the origin of keroliteformation. The dip and strike of kerolite and polygonal ser-pentine fractures show that their mean orientations are close(Fig. 11). Moreover, kerolite occurrences are locally striated(Cluzel and Vigier 2008). Cracking of inherited fracturescould be the result of their tectonic reactivation under a com-patible stress field. A recent study shows that in the case ofweak veins, typically the case of serpentine and kerolite veins,the reactivation of the fracture occurred even if the vein ori-entation is highly unfavorable relative to the stress field (Virgoet al. 2014). As shown previously, there are also fracturesfilled by kerolite but with no evidence of polygonal serpen-tine. These fractures differ, at least in part, in orientation fromthe mean common orientation of polygonal serpentine andkerolite fractures and could be considered newly formed.However, it appears difficult to provide a kinematic analysisfor the kerolite fractures given that kerolite do not providegood kinematic criteria.

In addition of the tectonic influence, the effect of localfluid overpressure cannot be excluded. Indeed, occurrencesof silica breccia are consistent with fluid-assisted deforma-tion following Sibson’s model (Sibson 1986). Taking a rea-sonable value of tensile strength for the wall rock (50 bars,Etheridge 1983), Sibson considers that for a depth

z < 0.5 km, no implosion of fracture sidewalls should occurand there is no hydrothermal precipitation in extensionalveins. As a deeper depth is required for brecciation, it ispermissive that the processes occurred at more than 0.5kmdepth. The tectonic context favored fracture opening andsubsequent upward fluid migration, which in turn enhancedinherited fracture opening and brecciation under local fluidoverpressures.

Concerning the breccia samples, the following conclusionsmay be made:

– The boundary fractal dimension of the clasts, defined asDr = 1.034, falls in the range of values typical for a phys-ical (mechanical) breccia formed when the amount ofstress exceeds the brittle resistance of the material(Griffith mode) (Jébrak 1997).

– The variation in fragment size approximates a power lawdistribution with corresponding value of fractal dimen-sion Ds = 1.144. Such a low PSD value is often observedin hydraulic breccias or in those where transportationcaused a size classification (Jébrak 1992; Jébrak 1997;Blenkinsop 1991).

– The brecciation process generated in situ fragmentationwith a jigsaw puzzle pattern and absence of significantrotation of the fragments, which are typical features ofhydraulic fracturing (Jébrak 1997).

– The matrix of the breccia is homogeneous and representsonly two co-precipitated minerals: microcrystallinequartz and microparticles of goethite. The lack of severalmineral assemblages as well as lack of banding is usualfor fluid-assisted brecciation, since (i) the pressure fluc-tuation provokes rapid mineral precipitation and (ii) hy-draulic fracturing has typically a shorter duration in com-parison with tectonic activity, throughout which the min-eral deposition conditions may vary (Jébrak 1997).

Thus, breccia samples have all the typical features of hydrau-lic fracturing (Jébrak 1997). This process is generally caused bydifferences of pressure between aquifers and the sudden con-nection between aquifers during seismic events. The increase influid pressure may have several origins, including a decrease infault permeability (due to mineral deposition or fault slip). Thebreccias occur at specific stages of the vein filling, as attested bythe different types of hydraulic breccias: (i) breccia with serpen-tine cement (polygonal serpentine) around serpentine clasts(lizardite), indicating hydrothermal activity started within thestability field of serpentine; (ii) breccia with kerolite cementsand serpentine (all polytypes) clasts; and, finally, (iii) brecciawith a reddish microcrystalline quartz cement and clasts of alltypes (serpentine, kerolite). The time evolution of the cementbreccia mineralogy always follows the same sequence, which isalso similar to the fracture cement chronology: (i) serpentine,(ii) kerolite, and (iii) red-brown quartz.

Miner Deposita

Page 114: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Conditions of red-brown quartz precipitation

The inferred model of Ni-kerolite + red quartz formation relieson the following observations and considerations: (i) Ni-Mgkerolite is intimately associated in space and time with redquartz indicating similar conditions during fracturing and de-position and (ii) trivalent iron in iron oxyhydroxides that co-precipitate with quartz is poorly soluble under oxidizing con-ditions. According to Stumm and Morgan (1996), the rate lawfor oxidation of Fe2+ at 20 °C is first order with respect to bothFe2+ and fO2 and second order for [OH−]. Thus, the time forcomplete oxidation of iron in an aerated solution is a matter ofhours under the pH conditions typical of laterite (waters equil-ibrated with atmospheric CO2 (pH 5.6) and of seconds undersaprolite conditions (pH ranging from 7 to 9). The only pro-cess accounting for a transport of trivalent iron could be me-chanical downward migration of Fe nanoparticles from theabove aquifers (Fig. 12). Such a transport might have beenpartially possible by the formation of fractures due to tectonicactivity and downward migration of the ferrihydrite colloid.Alternately, reducing conditions favor iron under the +2 oxi-dation state in anoxic solution and subsequent iron transport.Such reducing conditions occur below the water table, in thedeeper parts of the aquifers (Fig. 13). The co-precipitation ofred quartz needs, therefore, the iron oxidation and thesyncrystallization of quartz (Fig. 12). Such a process is com-monly used in industry to obtain silica-iron hydroxide co-precipitation (Cortez Fernandes et al. 2013; Dyer et al.2012). Silica precipitation has been shown to be strongly en-hanced by adsorption processes on freshly precipitated ferri-hydrite, serving as nuclei in oversaturated solutions, with

maximum sorption attained at pH 7.25–9.5 (Davis et al.2001, 2002; Swedlund et al. 2011). Under the typical pHvalues of waters at equilibrium with ultrabasic rocks, i.e.,around 9 to 10, it is likely that silica is mostly polymeric(Dyer et al. 2012) (Fig. 12) and, therefore, should stronglybind to iron. Thus, the formation of numerous germs of ironoxyhydroxides causes the co-precipitation of SiO2. The over-saturation of the fluid with respect to silica is also favored by atemperature drop and/or a pH decrease, since silica and quartzsolubilities decrease with decreasing temperature and pH. Asthe red-brown microcrystalline quartz represents a singleevent and as all other quartz types are free of oxyhydroxideparticles, especially those found close to surface in the silici-fied bedrocks, the hypothesis of upward migration of reducedwaters seems the most probable for this stage, in contrast tothe other quartz stages.

Crystallization temperatures of red-brown quartz

During the Ni stage, the late microcrystalline red quartz is adistinctive feature of the kerolite fractures and hydraulic brec-cia. Microcrystalline quartz always postdates kerolite precip-itation, but it seems to be related to the same cycle of fracturecracking and sealing, being geometrically associated tokerolite, and containing variable but significant amount ofNi. The oxygen isotopic composition of microcrystallinequartz from Koniambo ranges from 23.9 to 27.6 ‰(Quesnel 2015; Quesnel et al. 2016a). These values corre-spond to calculated temperatures in the range of ~50 to95 °C (using δ18OFluide = −2.5 ‰; Quesnel et al. 2016a).Therefore, talc-like formed at a temperature between that ofthe latest serpentines and a minimum temperature of about50 °C for quartz. Suchmedium-range temperature was alreadyproposed by Ducloux et al. (1993) for Bou Azzer kerolitealthough their temperature range is poorly constrained.

Fig. 12 Model proposed for the precipitation of the goethitemicrocrystalline quartz. Left-hand side: hypothetic downward transportof trivalent iron as nanoparticles. Right-hand side: upward circulation ofreduced water transporting divalent iron and co-precipitation of silica andiron hydroxides during mixing with oxidizing surface waters

Fig. 13 Conceptual modeling of the hydrothermal cell at the origin of thevein infillings

Miner Deposita

Page 115: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

A model of low-to-medium-temperature fluid circulationfor the formation of the Ni silicate veins

Since the microcrystalline quartz ends the main stage of for-mation of Ni-Mg kerolite, this may indicate that a great part ofNi silicate fracture fillings formed during a series of hydrolicfracturing, which cracks previous fractures, opens new ones,or causes brecciation. Fractures are well oriented and formsubparallel sets strikingN-S to NE-SW.Upward hydrothermalfluid migration of moderate temperature is inferred for thesestages of fracture sealing (Fig. 13). The successive precipita-tion of talc-like followed by microcrystalline quartz isinterpreted to occur when upward ascending fluids encountermore oxidizing waters near the surface. Such a model ofmedium-temperature upward circulation of reduced waters,as well as the high pressure necessary for hydraulic fracturingor brecciation in relation with tectonics, constitutes an alter-native to classic metallogenic models developed for NewCaledonian deposits. Thus, a unique downward infiltrationof cold oxidizing meteoric water under hydrostatic or infra-hydrostatic pressure cannot explain all the features of theKoniambo deposit. In the proposed model, fluids are consid-ered to transfer quickly from surface (meteoric water), areheated at depth, and then advect within the nappe. The fluidsare interpreted to mix during their upward migration withmeteoric, oxidized water potentially charged in nickel fromthe surface water-rock interactions. Therefore, the nickel isunlikely to be dissolved from the deepest part.

Conclusion

Most of the veins sealed by Ni-Mg kerolite and quartz-hematitecorrespond to stages of cracking and sealing of preexisting frac-tures probably formed by fluid overpressure. The tectonic re-gime at that stage was compatible with the reopening of earlierfractures previously cemented by lizardite, which was then par-tially to entirely converted to polygonal serpentine.

The systematic occurrence of a main Ni silicate stagemarked by the presence of Ni-Mg kerolite, formed in be-tween the latest stage of serpentine and the oxyhydroxidemicrocrystalline stage, provides a strong argument in favorof a moderate-temperature hydrothermal event. The micro-crystalline quartz, formed under temperatures of 50–95 °Cduring the last stage of the hydrothermal system, providesthus a good estimate of the minimum temperature condi-tions driving the preceding stages. It is also a good markerof divalent iron migration in a reduced environment, as itappears unlikely that iron could have been transported fromsurface as nanoparticles. Such Fe-hydroxides could havebeen formed by the oxidation of Fe2+, transported by re-duced water from underlying aquifers.

The sequence of fracture filling is therefore interpreted asthe result of a hydrothermal system evolution frommedium tolow temperature under fluctuating fluid pressure and depth>0.5 km. Data from Koniambo massif Ni-mineralized frac-tures allow us to propose a new genetic model that includesfluid advection driven with tectonic stress. This model is dif-ferent from the supergene model generally accepted for theformation of Ni silicate ores. The clusters of fractures closelylinked to reopening of former serpentine fractures indicate thata Ni preconcentration occurred in relation with tectonics and/or hydraulic processes. Although frequently described as theresult of a single downward redistribution of Ni and Mgleached in the upper part of the regolith, the Ni silicate veinsappear to have formed as a result upward of fluid flow.

At the scale of mineralized massif, the succession of fluidflow events produced local enrichment in nickel to account forpart of the heterogeneity in the distribution of nickel inpresent-day regolith. Fracture fillings are affected then bydissolution-recrystallization processes within the water tablemovement zone, a process yielding to the formation of target-like ores, and finally by the downward migration of theoxidation-dissolution front which forms saprolitic ore. Thepristine Ni-rich fracture distribution therefore controls partial-ly the geometry of the secondary Ni distribution. The present-day distribution of minerals appears thus as the result of com-plex superimposed processes and not only as the result ofsimple downward migration of a nickel front during the mainstage of laterite formation.

Acknowledgments This work has been carried out thanks to the finan-cial and technical help from Koniambo S.A.S. and Labex Ressources 21(supported by the French National Research Agency through the nationalprogram BInvestissements d’avenir,^ reference ANR-10-LABX-21-LABEX RESSOURCES 21), with analytical contributions from theUMR GeoRessources No. 7356 laboratory platforms. The field workwas introduced by a first 1-day visit at Koniambo in 2011 by C.Couteau and E. Fritsch (field trip organized as part of a CNRT project(French National Centre for Technological Research: BNickel and itsenvironment^)), and consisted in 2-week stays both in 2012 and 2013financed by Koniambo within the framework of the B. Quesnel’s PhDthesis. This work was supervised in the field by C. Couteau and MaximeDrouillet from Koniambo S.A.S. Marie-Camille Caumon is acknowl-edged for her help during RAMAN analyses. The authors are gratefulto M. Jébrak, E. Ramanaidou, and B. Orberger for their detailed andconstructive reviews.

References

Allmendinger RW, Cardozo NC, Fisher D (2013) Structural geologyalgorithms: vectors & tensors. Cambridge University Press,Cambridge, 289 pp

Auzende A, Daniel I, Reynard B, Lemaire C, Guyot F (2004) Highpressure behaviour of serpentine minerals: a Raman spectroscopicstudy. Phys Chem Miner 31(5):269–277

Baronnet A, Devouard B (1996) Topology and crystal growth of naturalchrysotile and polygonal serpentine. J Cryst Growth 166:952–960

Miner Deposita

Page 116: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Bérubé D, Jébrak M (1999) High precision boundary fractal analysis forshape characterization. Comput Geosci 25:1059–1071

Blenkinsop TG (1991) Cataclasis and processes of particle size reduction.Pure Appl Geophys 136:59–86

Bonnet E, Bour O, Odling NE, Davy P, Main I, Cowie P, Berkowitz B(2001) Scaling of fracture systems in geological media. Rev.Geophys 39(3):347–383

Brindley GW, Hang PT (1973) The nature of garnierites—I structures,chemical compositions and color characteristics. Clay Clay Miner21:27–40

Brindley GW, Wan HM (1975) Compositions, structures, and thermalbehaviour of nickel-containing minerals in the lizardite–nepouiteseries. Am Mineral 60:863–871

Brindley GW, Bish D, Wan HM (1979) Compositions, structures, andproperties of nickel-containing minerals in the kerolite–pimelite se-ries. Am Mineral 64:615–625

Butt CRM, Cluzel D (2013) Nickel laterite ore deposits: weatheredserpentinites. Elements 9(2):123–128

Cathelineau M, Caumon MC, Massei F, Brie D, Harlaux M (2015b)Raman spectra of Ni-Mg kerolite: effect of Ni-Mg substitution onO-H stretching vibrations. Proceedings GEORaman 2014 (11th), StLouis (USA). J Raman Spectrosc 46(10):933–940

CathelineauM,Quesnel B, Gautier P, Boulvais P, Couteau C, Drouillet M(2015a) Nickel dispersion and enrichment at the bottom of the reg-olith: formation of pimelite target-like ores in rock block joints(Koniambo Ni deposit, New Caledonia). Mineral Deposita 51:271–282

Chardon D, Chevillotte V (2006) Morphotectonic evolution of the NewCaledonia ridge (Pacific Southwest) from post-obductiontectonosedimentary record. Tectonophysics 420:473–491.doi:10.1016/j.tecto.2006.04.004

Chen TT, Dutrizac JE, Krause E, Osborne R (2004) Mineralogical char-acterization of nickel laterites from New Caledonia and Indonesia.In: International Laterite Nickel Symposium (as held during the2004 TMS annual meeting). p. 79-99

Cluzel D, Vigier B (2008) Syntectonic mobility of supergene nickel oresof New Caledonia (Southwest Pacific). Evidence from garnieriteveins and faulted regolith. Resour Geol 58:161–170. doi:10.1111/j.1751-3928.2008.00053.x

Cluzel D, Aitchison JC, Picard C (2001) Tectonic accretion and under-plating of mafic terranes in the Late Eocene intraoceanic fore-arc ofNew Caledonia (Southwest Pacific): geodynamic implications.Tectonophysics 340(1–2):23–59

Cluzel D, Jourdan F, Meffre S, Maurizot P, Lesimple S (2012) The meta-morphic sole of New Caledonia ophiolite: 40Ar/39Ar, U-Pb, andgeochemical evidence for subduction inception at spreading ridge.Tectonics 31, TC3016, doi: 10.1029/2011TC003085.

Cortez Fernandes MT, Rivero Garcia RB, Paula Leite CA, Kawachi EY(2013) The competing effect of ammonia in the synthesis of ironoxide/silica nanoparticles in microemulsion/sol–gel system.Colloids and Surfaces A: Physicochem Eng Aspects 422:136–142

de Faria DLA, Venâncio Silva S, de Oliveira MT (1997) Raman spec-troscopy of some iron oxides and oxyhydroxides. J RamanSpectrosc 28:873–878

Davis CC, Chen H, Edwards M (2002) Modeling silica sorption to ironhydroxide. Environmental Science & Technology 36(4):582–587.doi: 10.1021/es010996t

Davis CC, KnockeWF, EdwardsM (2001) Implications of aqueous silicasorption to iron hydroxide: mobilization of iron colloids and inter-ference with sorption of arsenate and humic substances.Environmental Science & Technology 35(15):3158–3162. doi:10.1021/es0018421

Dublet G, Juillot F, Morin G, Fritsch E, Fandeur D, Brown GE (2015)Goethite aging explains Ni depletion in upper units of ultramaficlateritic ores from New Caledonia. Geochim Cosmochim Acta160:1–15

Dublet G, Juillot F, Morin G, Fritsch E, Fandeur D, Ona-Nguema G,Brown GE Jr (2012) Ni speciation in a New Caledonian lateriticregolith: a quantitative X-ray absorption spectroscopy investigation.Geochim Cosmochim Acta 95:119–133

Ducloux J, Boukili H, Decarreau A, Petit S, Perruchot A, Pradel P (1993)Un gîte hydrothermal de garniérites: l’exemple de Bou Azzer,Maroc. Eur J Mineral 5:1205–1215

Dyer LG, Richmond WR, Fawell PD (2012) Simulation of iron oxide/silica precipitation in the paragoethite process for the removal of ironfrom acidic zinc leach solutions. Hydrometallurgy 119–120:47–54

Etheridge MA (1983) Differential stress magnitudes during regional de-formation and metamorphism: upper bound imposed by tensile frac-turing. Geology 11:231–1234

Fandeur D (2009) Géochimie et cristallochimie du chrome au cours del’altération de roches ultrabasiques en Nouvelle-Calédonie (Massifdu Koniambo). Dissertation, Université Paris Diderot–IPGP

Freyssinet P, Butt CRM,Morris RC (2005) Ore-forming processes relatedto lateritic weathering. Econ Geol: 681–722. 100th Anniversary

Fritsch EM, Juillot FA, Dublet GA, Fonteneau L, Fandeur D, Martin E,Caner L, Auzende A-L, GraubyO, Beaufort D (2016) An alternativemodel for the formation of hydrous Mg/Ni layer silicates(Bdeweylite^/Bgarnierite^) in faulted peridotites of New Caledonia:I. Texture and mineralogy of a paragenetic succession of silicateinfillings. Eur J Mineral 28(2):295–311

Golightly JP (2010) Progress in understanding the evolution of nickellaterite. Society of Economic Geologists, Inc. Special Publication15:451–485

Jébrak M (1992) Les textures intra-filoniennes, marqueurs des conditionshydrauliques et tectoniques. Chron Rech Min 506:55–65

Jébrak M (1997) Hydrothermal breccias in vein-type ore deposits: a re-view of mechanisms, morphology and size distribution. Ore GeolRev 12:111–134

Karperien A (1999–2013) FracLac for ImageJ http://rsb.info.nih.gov/ij/plugins/fraclac/FLHelp/Introduction.htm

Lach P, Mercadier J, Dubessy J, Boiron MC, Cuney M (2013) In-situquantitative measurement of rare earth elements in uranium oxidesby laser ablation-inductively coupled plasma-mass spectrometry.Geostandards and Geoanalytical Research 37:277–296

Leguéré J (1976) Des correlations entre la tectonique cassante etl’altération supergene des peridotites de Nouvelle-Calédonie.Dissertation, Université du Languedoc, Montpellier, France

Leisen M, Dubessy J, Boiron MC, Lach P (2012) Improvement of thedetermination of element concentrations in quartz-hosted fluid in-clusions by LA-ICP-MS and Pitzer thermodynamic modeling of icemelting temperature. Geochim Cosmochim Acta 90:110–125

Lemaire C (2000) Application des spectroscopies vibrationelles à la de-tection d’amiante dans les matériaux et à l’étude des serpentines.Dissertation, Université de Paris 07.

Longerich HP, Jackson SE, Günther D (1996) Laser ablation inductivelycoupled plasma mass spectrometry transient signal data acquisitionand analyte concentration calculation. J Anal At Spectrom 11:899–904

Manceau A, Calas G (1985) Heterogeneous distribution of nickel in hy-drous silicates from New Caledonian ore deposits. Am Mineral 70:549–558

Manceau A, Calas G (1986) Nickel-bearing clay minerals: II.Intracrystalline distribution of nickel: an X-ray absorption study.Clay Miner 21:341–360

Maurizot, P, Lafoy, Y, Poupée M (2002) Cartographie des formationssuperficielles et des aléas mouvements de terrain en Nouvelle-Calédonie, Zone du Koniambo: Bureau de RecherchesGéologiques et Minières, Public Report RP51624-FR, 45 p

Maurizot P, Vendé-Leclerc M, Collot J (2009) Carte géologique de laNouvelle-Calédonie au 1/500.000. Service Géologique deNouvelle-Calédonie/DIMENC-BRGM

Miner Deposita

Page 117: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Paquette JL, Cluzel D (2007) U–Pb zircon dating of post-obduction vol-canic-arc granitoids and a granulite-facies xenolith from NewCaledonia. Inference on Southwest Pacific geodynamic models.Int J Earth Sci 96:613–622. doi:10.1007/s00531-006-0127-1

Pearce NJG, Perkins WK, Westgate JA, Gorton MP, Jackson SE, Neal CR,ChenerySP (1997)Acompilationofnewandpublishedmajor and traceelement data for NIST SRM 610 and NIST SRM 612 glass referencematerials. Geostandards andGeoanalytical Research 21:101–114

Quesnel B (2015) Altération supergène, circulation des fluides et defor-mation interne du massif de Koniambo, Nouvelle-Calédonie:Implication sur les gisements nickélifères latéritiques. Dissertation,Rennes Univ., 276 p

Quesnel B, Boulvais P, Gautier P, Cathelineau M, John CM, Dierick M,Agrinier P, Drouillet M (2016a) Paired stable isotope (O, C) andclumped isotope thermometry of magnesite and silica veins in theNew Caledonia Peridotite Nappe. Geochim Cosmochim Acta 183:234–249

Quesnel B, Gautier P, CathelineauM, Boulvais P, Couteau C, Drouillet M(2016b) The internal deformation of the Peridotite Nappe of NewCaledonia: a structural study of serpentine-bearing faults and shearzones in the Koniambo Massif. J Struct Geol 85:51–67

Rasband WS (1997–2014) ImageJ. US National Institutes of Health,Bethesda, Maryland, USA, http://imagej.nih.gov/ij/

Sevin B, Ricordel-PrognonC, Quesnel F, Cluzel D, Lesimple S,MaurizotP (2012) First palaeomagnetic dating of ferricrete in NewCaledonia:new insight on themorphogenesis and palaeoweathering of BGrandeTerre^. Terra Nov. 24(1):77–85

Sibson RH (1986) Brecciation processes in fault zones. Pure ApplGeophys 124:159–175

Stumm W, Morgan JJ (1996) Aquatic chemistry, 3rd edn. Wiley andSons, New York, 1022 pp

Swedlund PJ, Sivaloganathan S, Miskelly GM, Waterhouse GIN (2011)Assessing the role of silicate polymerization on metal oxyhydroxidesurfaces using X-ray photoelectron spectroscopy. Chem Geol 285:62–69

Trescases JJ (1975) L’évolution géochimique supergène des rochesultrabasiques en zone tropicale. Mem ORSTOM, 78, 259 p

Turcotte DL (1986) Fractal and fragmentation. J Geophys Res 91:1921–1926

USGS (US Geological Survey) (2016) Mineral commodity summaries,J a n u a r y 2 0 1 6 . h t t p : / / m i n e r a l s . u s g s .gov/minerals/pubs/commodity/nickel/mcs-2016-nicke.pdf

Villanova-de-Benavent C, Proenza JA, Galí S, García-Casco A, Tauler E,Lewis JF, Longo F (2014) Garnierites and garnierites: textures, min-eralogy and geochemistry of garnierites in the Falcondo Ni-lateritedeposit, Dominican Republic. Ore Geol Rev 58:91–109

Virgo S, Abe S, Uria JL (2014) The evolution of crack seal vein andfracture networks in an evolving stress field: insights from discreteelement models of fracture sealing. J Geophys Res 119-12:8708–8727

Wells MA, Ramanaidou ER, Verrall M, Tessarolo C (2009) Mineralogyand chemical chemistry of Bgarnierites^ in the Goro lateritic nickeldeposit, New Caledonia. Eur J Mineral 21:467–483

Wells MA, Ramanaidou ER (2015) Variability of the Ni content in goe-thite for New Caledonian lateritic nickel deposits. In: Proceedings ofthe 13th Biennal SGA (Society for Geology Applied to MineralDeposits) meeting, Nancy, 24-27 August 2015

XStrata (2012) Mineral resources and reserves at 31 Dec. 2012. Ann Rep50

Miner Deposita

Page 118: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 119: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genesis and mineralogy of Ni ores

Myagkiy A., Cathelineau M., Golfier F., Truche L., Quesnel B., Drouillet, M., (2016b).

Defining mechanisms of brecciation in Ni-silicate bearing fractures (Koniambo, New Caledonia).

SGA Conference Nancy 2015 Proceedings, vol. 3, pp . 1181-1185.

- 8 -

Page 120: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Defining mechanisms of brecciation in Ni-silicate bearing fractures (Koniambo, New Caledonia) Myagkiy A, Cathelineau M, Golfier F, Truche L Université de Lorraine, CNRS, CREGU, GeoRessources lab., F-54518, Vandoeuvre-lès-Nancy, France Quesnel B Géosciences Rennes, Université Rennes 1, UMR 6118 CNRS, 35042 Rennes Cedex, France Drouillet M Service géologique, Koniambo Nickel SAS, 98883 Voh, Nouvelle Calédonie Abstract The origin of the breccias that were found in faults from bedrocks of New Caledonia remains poorly known. In this way, study of the breccias that affect Ni-silicate ore fracture fillings is of interest and appear to be a key point to understand the early Ni-silicate ore formation processes.

Microcrystalline quartz-hematite cements clasts of serpentine and Ni talc-like in fault breccia from Koniambo open pits. This breccia was investigated in the current study. Here we quantitatively analyze the dimensions and shape of clasts along with a matrix composition and dilation in order to find out what process governed brecciation. Breccia clast morphology and size distribution are quantified using fractal analysis techniques. The result of these analyses indicates that investigated breccia was formed due to hydraulic fracturing, which is related to an increase in fluid pressure. Keywords: breccia, morphology, particle size distribution, hydrofracturing 1 Introduction Laterite nickel-ore formation in New Caledonia is classically supposed to be controlled only by the weathering processes, that lead to residual soils in subsurface (laterite) and subsequent downward migration of Si, Mg and Ni (Trescases, 1975, Genna et al., 2005 ). However, this conceptual model fails to explain some recent mineralogical and structural observations (Quesnel et al. 2013). In particular, none of supergene models succeed to describe the heterogeneity of mineralizing distributions that could have been enhanced by several syntectonic pre-enrichment cycles. Quesnel et al (2013) and Cluzel & Vigier (2008) have respectively shown that at least a part of magnesite veins and a part of the Ni talc-like veins formed syn-tectonically.

To date, there is no study focused on the relationship between silica veins formation and deformation, whereas i) the silica veins occur as faults and fracture fillings ii) they often consist of complex textural association of several types of silica and iii) they sometimes occur as breccia where silica cements clasts of serpentine and Ni talc-like. Such breccia may have

been formed by downward or upward fluid flow. An eventual upward flow in New Caledonia ophiolite could be related to hydrothermal fluid circulations. Evidence for present-day low-temperature (50°C), high-pH hydrothermal system in New Caledonia has been previously reported by Launay and Fontes (1985). Monnin et al. (2014) show that these hydrothermal systems in New Caledonia are fed by meteoric waters that percolate through the densely fractured peridotites, get significantly higher pH due to water-rock interactions and then discharge at springs in the lagoon where they mix with seawater. In any case, regardless their flow directions, these circulations could have caused pressure-induced fracturing in the ophiolite if the fluid pressure exceeds the rock mechanical resistance.

According to review of the main brecciation processes occurring in hydrothermal vein-type deposits (Jébrak 1997), there are eight main mechanisms of brecciation: tectonic comminution, wear abrasion, two types of fluid-assisted brecciation (hydraulic and critical), volume expansion or reduction, impact and collapse. Each of these mechanisms can be distinguished by carrying out quantitative analyses of the breccia.

In this study, four parameters were used to describe the breccia: (1) clast morphology, (2) particle size distribution, (3) composition of matrix and (4) dilation ratio. The clast morphology is related to the complexity of shape of individual clasts while particle size distribution was applied to all plurality of clusters. Fractal dimension of fragment morphology was performed using the box-counting method, developed specifically for computerized fractal analysis and implemented in the plugin “FracLac” (Karperien 1999-2013). Fractal dimension of particle size distribution was obtained using a cumulative distribution curve of the fragment sizes. Processing was carried out on Digital Microscope VHX-2000 (© Keyence) whereby the clusters and their geometrical sizes were identified. The dilation ratio was identified as the volume ratio of matrix and clast. These quantitative analyses were then compared with results of review of the main brecciation processes (Jébrak 1997). 2 Description of the breccia

Page 121: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Following oligomictic breccia (Figure 1), found in some Koniambo fractures from the bedrock and saprolite, has been investigated in order to determine its brecciation process. The breccia comprises angular serpentine clasts (lizardite, partially to totally converted to polygonal serpentine, as defined by Baronnet and Devouard, 1996) and bluish-green Ni-Mg talc-like phase clasts (also called kerolite), within a reddish-brown microcrystalline quartz-hematite matrix. The mineralogy of the clasts and matrix was analyzed by Raman spectroscopy. The texture of the breccia is matrix-supported with a high value of dilation. The size of the fragments ranges from few hundred to few thousand micrometers.

Figure 1. Photographs of investigated breccia, made by Digital Microscope VHX-2000 using the option “phase assemblage”. Clasts of dark colour represent lizardite while those of light green-colour are identified as serpentine-like, and talc-like when of bluish-green color; the matrix is composed of microcrystalline quartz hosting dispersed fine-grained hematite particles.

Jigsaw textures are locally recognized at grain scale (Figure 2). Fragments still match along adjacent surfaces and show that they are parts of once unbroken clast. Angular shape of fragments implies that movements were not sufficient to round them. It should be noted that local jigsaw textures are typically formed by fluid-assisted brecciation (Jébrak 1997).

Figure 2. Photographs of clasts, which were disjointed during brecciation process.

3 Fractal analysis In current study two fractal dimensions Dr (the complexity of the clasts) and Ds (the particle size distribution) were used to describe the breccia. In order to define the morphology of the fragments, a “FracLac” (Karperien 1999-2013) plugin for ImageJ (Rasband 1997-2014) had been used. The method of fractal analysis for calculation Dr was the box-counting method. Averaging adjoining pixels to form a new image with coarser base elements progressively coarsens the representation of the object (Bérubé and Jébrak 1999). Box-counting was performed on one-pixel outlines of clasts that were generated using ImageJ software and “automatic measurements” tool of Digital Microscope VHX-2000. An example of outline that was used is shown on Figure 3.

Average boundary fractal dimension Dr obtained for clasts of breccia is Dr = 1.034±0.029 and means that clasts have an angular nearly Euclidean shape. Based on these first results, it can be concluded that the breccia was exposed to physical (mechanical) brecciation and occurred when the amount of stress exceeded the brittle resistance of the material (Griffith mode) (Jébrak 1997).

Page 122: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Figure 3. The examples of outline of clasts, obtained by ImageJ and Digital Microscope VHX-2000. Without scale as it does not affect the measurement of complexity of the clasts. 4 Particle size distribution For convenience of processing, the photo of breccia (Figure 1) was converted into black and white (Figure 4). 209 clasts were allocated for characterization of size distribution. Their size range is 122.04 µm<r <4744.64 µm, where r is the equivalent spherical radius of clasts. “r” values were calculated as square roots of the clasts areas in order to be in accordance with the method of Turcotte (1986).

According to Turcotte (1986) if the number-size distribution of objects satisfies the condition N∼r (-Ds) (1) (where “N” is the number of objects with a characteristic linear dimension greater than “r”), then a fractal is defined with fractal dimension Ds.

Built-in tool of Digital Microscope VHX-2000 was used to measure the clast areas.

Figure 4. Conversion of the photo into black and white in order to highlight the clasts (black) from the reddish brown silica matrix (white).

The particle size distribution (PSD) of 209 defined clasts was then described by a power low relationship (1). The slope of best-fit line on a log-log plot of N(>r) vs. r (Figure 5) is equivalent to the fractal dimension, Ds, of the particle size distribution (Turcotte 1986) and found to be 1.144. Particles larger than 2834 µm deviate from the power law relationship because of censoring effects (Bonnet et al 2001). These effects imply that fragments with sizes, comparable to the sampled region, have a tendency to deviate from power law distribution.

Since PSD is a function of the energy input during

breccia formation (Jébrak 1997), the low value of Ds indicates low energy of brecciation process. A dilation ratio of investigated breccia is found to be high: the matrix is very abundant and represents 84% of area. It is worth noting that the matrix of breccia is homogeneous and represents only two co-precipitated minerals: microcrystalline quartz and microparticles of hematite. The lack of several mineral assemblages, which are typical for brecciation due to tectonic activity, is usual for fluid-assisted brecciation since the pressure fluctuations provoke a rapid mineral precipitation (Jébrak 1997).

Figure 5. Fractal dimension of clast size distribution, Ds, of investigated breccia. 5 Evidence of fluid-assisted fracturing As a summary, the following observations have been made:

- clasts from the studied breccias have an angular nearly Euclidean shape with corresponding value of fractal dimension of clast morphology Dr=1.034. This falls in the range for a physical (mechanical) breccia formed when the amount of stress exceeded the brittle resistance of the material.

- The variation in fragment sizes in the investigated breccia approximates a power-law distribution (Figure 5) with corresponding value Ds =1.144. Low values of fractal dimension of clast size distributions are commonly associated with fluid-assisted processes of brecciation (Jébrak 1992, Blenkinsop 1991).

- Local jigsaw textures, abundance of quartz matrix and high dilation ratio were noted along with simple filling material of matrix consisting of co-precipitated microcrystalline quartz and microparticles of hematite.

All of these observations supported by the results of fractal analysis are evidences of fluid-assisted fracturing (Jébrak 1997).

5 Discussion and Conclusions

The brecciation process occurs at specific stages of long-lived hydrothermal and tectonic activity, as different types of hydraulic breccias were found: i) breccia with serpentine cement (polygonal serpentine) around serpentine clasts (lizardite), indicating a starting

Page 123: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

of the hydrothermal activity within the stability field of serpentine, ii) breccia with kerolite cements with serpentine (all polytypes) clasts, and finally, iii) breccia with a reddish microcrystalline quartz cement and clasts of all types (serpentine, kerolite). The time evolution of the cement mineralogy always represents the same sequence: i) serpentine, ii) kerolite, iii) reddish quartz.

The third main cement, i.e. microcrystalline quartz hosting micro-particles of hematite is a very distinctive feature of the hydraulic breccia. Thus, as trivalent iron is almost immobile under oxidizing conditions, no iron can be transported by the penetration of downward oxidising meteoric water. At variance, iron could have been transported as Fe 2+ by deep-seated fluids likely coming from the bottom of the ophiolite. The fluids at the origin of the hydraulic breccias were therefore reduced fluids and then underwent oxidation during the hydraulic brecciation, and subsequently precipitated hematite. Thus, as soon as that fluid, moving upward, has come in contact with a meteoric one, more oxidizing and slightly cooler, oversaturation with respect to quartz and iron oxidation happened. Geochemical modelling is in progress in order to understand these mineral sequences.

Consequences of these observations are rather important for the inferred models of ore fluid circulations. Since the reddish micro-crystalline quartz ends the main stage of formation of bluish-green kerolite, this may indicate that a great part of Ni-silicate fracture infillings were formed during a series of fluid-assisted fracturing and brecciation process. The latter implies upward hydrothermal fluid migration. Such a model is at variance of metallogenic models developed for New Caledonian deposits and dealing with supergene processes. Acknowledgements Authors sincerely thank the LabEx RESSOURCES21 for financial support and geological service of Koniambo Nickel SAS for close collaboration. References Baronnet, A., Devouard B., 1996. Topology and cristal growth of

natural chrysotile and polygonal serpentine. Journal of Crystal Growth, Vol. 166, pp. 952-960.

Bérubé D, Jébrak M (1999) High precision boundary fractal analysis for shape characterization. Computers & Geosciences 25:1059-1071.

Blenkinsop TG (1991). Cataclasis and processes of particle size reduction. Pure Appl. Geophys. 136: 59–86.

Bonnet E, Bour O, Odling NE, Davy P, Main I, Cowie P and Berkowitz B (2001), Scaling of Fracture Systems in Geological Media, Rev. Geophys., 39(3), 347-383.

Cluzel D, Vigier B (2008) Syntectonic mobility of supergene nickel ores of New Caledonia (Southwest Pacific): Evidence from garnierite veins and faulted regolith. Resource Geology 58:161-170.

Genna A, Bailly L, Lafoy Y, Augé T (2005) Les karts latéritiques de Nouvelle Calédonie, Karstologia n°45-46, pp 19-28

Jébrak M (1992) Les textures intra-filoniennes, marqueurs des conditions hydrauliques et tectoniques. Chron. Rech. Min. 506, 55–65.

Jébrak M (1997) Hydrothermal breccias in vein-type ore deposits: A review of mechanisms, morphology and size distribution. Ore geology reviews 12:111-134.

Karperien A (1999-2013) FracLac for ImageJ. http://rsb.info.nih.gov/ij/plugins/fraclac/FLHelp/Introduction.htm Launay J, Fontes JC (1985) Les sources thermales de Prony

(Nouvelle Calédonie) et leurs précipités chimiques, exemple de formation de brucite primaire, Geologie de la France, 1, 83– 100.

Monnin C, Chavagnac V, Boulart C, Ménez B, Gérard M, Gérard E, Pisapia C, Quéméneur M, Erauso G, Postec A, Guentas- Dombrowski L, Payri C and Pelletier B (2014) Fluid Chemistry of the low temperature hyperalkaline hydrothermal system of Prony Bay (New Caledonia). Biogeosciences, 11, 5687–5706.

Quesnel B, Gautier P, Boulvais P, Cathelineau M, Maurizot P, Cluzel D, Ulrich M, Guillot S, Lesimple S, Couteau C (2013) Syn-tectonic, meteoric water-derived carbonation of the New Caledonia peridotite nappe. Geology 41:1063–1066.

Rasband WS (1997-2014) ImageJ. U. S. National Institutes of Health, Bethesda, Maryland, USA, http://imagej.nih.gov/ij/

Turcotte DL (1986) Fractal and fragmentation. Journal of Geophysical research 91:1921-1926.

Page 124: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 125: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genesis and mineralogy of Ni ores

Cathelineau M., Caumon, M-C, Massei F., Brie D. and Harlaux M. (2015)

Raman spectra of Ni–Mg kerolite: effect of Ni–Mg substitution on O–H stretching vibrations. J. Raman ; Spectroscopy,

DOI 10.1002/jrs.4746

- 9 -

Page 126: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Raman spectra of Ni–Mg kerolite: effect ofNi–Mg substitution on O–H stretching vibrationsMichel Cathelineau,a Marie-Camille Caumon,a* Frédéric Massei,a David Brieb

and Matthieu Harlauxa

The Ni-richmineral phases constituting the Ni-ores in New Caledonia are dominated inmost cases by kerolite–pimelite series, alsocalled talc-like minerals. Despite their economic and geologic interests, the crystallographic structures of these minerals are notfully understood. In order to improve the knowledge of their crystallographic structure, a set of natural talc-like minerals of vary-ing Ni/Mg ratio was selected. A combination of SEM, EMPA, TEM, and Raman analysis shows that the studiedmineral contains onlyone mineral phase identified as the kerolite–pimelite series. Although kerolite and pimelite are not currently recognized as dis-tinct minerals by the IMA, this study shows that these minerals exist as distinct minerals without any mixed layering with othermineral phases such as serpentine or other sheet silicates. The solid solution between the Ni and Mg endmembers was completewithout any gap, corresponding to a Ni–Mg substitution in the octahedral sheet. The Raman spectra in the OH stretching vibrationregion of representative samples covering the whole range of Ni–Mg substitution show a continuous and significant evolutionfrom the Mg to the Ni endmember. The signal was decomposed into nine Gaussian–Lorentzian functions, but the large overlap-ping of the peaks and the complexity of the band structure prevent any interpretation of the spectra from a structural point ofview. Finally, the problem is formalized as amultivariate curve resolution (MCR) which is solved using the Bayesian Positive SourceSeparation (BPSS) algorithm. This reveals that four possible arrangements of Ni and Mg in the octahedral sheet are encountered,and these arrangements are dependent on the Ni–Mg substitution rate. It also confirms that Raman microspectroscopy can bevery efficient in quick evaluation of the Ni content of the mineral. Copyright © 2015 John Wiley & Sons, Ltd.

Keywords: nickel; kerolite; Raman spectroscopy; chemometrics; OH

Introduction

The Ni-rich mineral phases constituting the so-called garnieriteNi-ores are dominated in most New Caledonia occurrences bytalc-like minerals, also called kerolite or pimelite depending onthe relative concentration of Mg or Ni.[1] Even if kerolite andpimelite are not formally recognized asmineral species by IMA, sev-eral occurrences are found in literature.[1–8] The Ni–Mg kerolite solidsolution from the two end-members, e.g. Mg-rich kerolite to Ni-richpimelite, was first described by Brindley et al.[2] Kerolite has a struc-ture close to the one of talc, with an interlayer distance of ~9.5Å butis characterized by an excess of Mg in octahedral site and a relativedeficit of Si in the tetrahedral site.[2,5] The H2O content is alsogreater than in talc. Brindley et al.[2] proposed an ideal chemical for-mula of (Ni,Mg)3 + x(Si4� y)O10(OH)2·nH2O, where x = 2y and n~1.Ni can substitute to Mg in octahedral sites, yielding to a completeseries from Mg–kerolite (noted Ni0) to pure Ni–kerolite (Ni100,pimelite) endmembers.[1,3,5]

These minerals have generally collomorph textures with succes-sive growing bands characterized by oscillatory changes in theNi/Mg ratio (e.g. Fig. 1). Wells et al.[1] and Villanova-de-Benaventet al.[5] showed that the entire range Ni0–Ni100 may be found in nat-ural occurrences. Previous studies[1–5] show that no mixed layeringbetween serpentines and talc-likeminerals occurs in the Ni–keroliteminerals, but also that the two kinds of minerals can coexist atmicrometric scale. In addition, the distribution of Ni in substitutionfor Mg in the octahedral site could be homogeneous or heteroge-neous because clusters of Ni at a scale of at least 30Å were sug-gested by Manceau and Calas.[4]

A preliminary Raman study of kerolite showed significantchanges in the Raman spectra as a function of the relative concen-tration in Ni and Mg.[7] Although several studies described the‘garnierite’ ores composed of serpentine, kerolite, and sepiolite bymeans of XRD, HR-TEM, chemical analyses by SEM, and electronmicroprobe analyses (EMPA),[1] almost no spectroscopy data onthe pure mineral phases is available on such minerals. Villanova-de-Benavent et al.[6] showed a Raman spectrum of pimelite in the100–1300 cm�1 range in order to distinguish the different Niphases in garnierites, but a correlation between the Ni content ofthe mineral phases and the position or intensities of the Ramanbands was not shown. Gerard andHerbillon[9] studied the evolutionof the infrared spectra of two kerolite samples. The spectra in therange of 3000–3800 cm�1 showed a complex structure at roomtemperature, with some sharp peaks superimposed onto wide,unstructured water bands. After thermal treatment at 800 °C, foursharp peaks appeared, correlated by the authors to the four possi-ble arrangements of Ni and Mg in the octahedral sites (Mg3, Mg2Ni,MgNi2, and Ni3) on the basis of the infrared study of Wilkins and

* Correspondence to: Marie-Camille Caumon, UMR GeoRessources, Faculté desSciences et Technologies, BP 70239, F-54 506 Vandœuvre-lès-Nancy, France.E-mail: [email protected]

a Université de Lorraine, CNRS, CREGU, GeoRessources laboratory, BP 70239,F-54506, Vandœuvre-lès-Nancy, France

b Université de Lorraine, CNRS, CRAN laboratory, BP 70239, F-54506, Vandœuvre-lès-Nancy, France

J. Raman Spectrosc. (2015) Copyright © 2015 John Wiley & Sons, Ltd.

Research article

Received: 30 September 2014 Revised: 19 May 2015 Accepted: 21 May 2015 Published online in Wiley Online Library

(wileyonlinelibrary.com) DOI 10.1002/jrs.4746

Page 127: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Ito[10] of Ni-enriched synthetic talcs. However, the weak differencein Ni content of their two kerolite samples and a possible structuraleffect of the thermal treatment on the crystallographic structure donot allow an unequivocal assignment of the peaks.The aim of the present study was to investigate the structure of

kerolite over the whole Mg–Ni substitution range by Ramanmicrospectroscopy. A particular focus was to examine the behaviorof the O–H stretching vibration bands with Ni content variations.Molecular Raman microspectroscopy was combined with atomicanalytical techniques (e.g. TEM-EDS and EMPA) to provide a corre-lation between chemical composition and structure of the mineral.Thus, a set of samples coveringmost of each 10% intervals betweenNi0 and Ni100 kerolite was selected. Raman spectra were acquiredon growing bands of polished sections previously analyzed byBSE-SEM and EMPA. The Raman signal was decomposed into ninepeaks, and also analyzed using the Bayesian Positive Source Sepa-ration (BPSS) chemometrics method. The results of decompositionand BPSS are compared and discussed in terms of mineral compo-sition and structure in light of chemical information.

Material and methods

Samples

The selected minerals occur as infillings of micro- to macro-fractures and generally exhibit chemical and color zonings becauseof oscillating Ni–Mg concentrations. Mg-rich kerolite end-member(Ni0) is whitish, Ni–Mg kerolite have bluish to greenish features,and Ni-rich pimelite (Ni100) ends the sequence with dark green toemerald color. The series were sampled in two open pits from thesites of Koniambo mine (open pit 207 and test pit from KoniamboSA) and Poro mine (SLN), New Caledonia. The samples were se-lected to cover a complete range of Ni contents between the twoendmembers of Ni–Mg kerolite composition.The samples were cut and polished in order to facilitate themap-

ping of mineral phases and chemical zonings by BSE-SEM and theselection of favorable areas to carry out Raman spectroscopy andchemical analysis by EMPA. Thus, a few zones characterized bychemical homogeneity on areas of more than 20 microns squarewere selected and mapped by SEM. Raman spectra were recordedat the exact point of chemical analysis by EMPA. Four samples wererequired to cover the whole range of Ni content as no sampleshowed a complete range of Ni–Mg substitution, although ratherlarge variations were recorded in a single sample.

SEM and EMPA

SEM images were obtained using the back-scattered electronsmode (BSE), in order to reveal the elemental zoning and changesin the mean atomic number (Z) using a JEOL J7600F SEM equippedwith SDD type electron dispersive spectrometer and Wave wave-length dispersive spectrometer (Oxford Instrument). Electronmicroprobe analyses (EMPA) of Si, Al, Fe, Mn, Mg, Co, Ni, and Nawere performed using a CAMECA SX100 instrument equipped witha WDS and calibrated using natural and synthetic oxides (albite (Si,Na), Al2O3 (Al), olivine (Mg), hematite (Fe), MnTiO3 (Mn), Co (Co), NiO(Ni)). The analytical conditions were a current of 12nA, an accelerat-ing voltage of 15 kV, and a counting time of 10 s (30 s to 60 s for Niin Ni-poor kerolite). The spatial resolution was of 1 micron. Total Feis presented as FeO. Structural formulae were calculated on the ba-sis of 22 negative charges per half unit cell, i.e. an O10(OH)2 basis. K,Na, Mn, and Cr were consistently found at levels near or below de-tection limits and, accordingly, were not included in structural for-mula calculations. All Si and any Al in the analyses were assumedto occupy only the tetrahedral sheet. The octahedral sheet was as-sumed to be occupied by Ni and Mg and the minor amounts of Feand/or Co found to be in some samples. The ratio Ni/(Ni +Mg) wascalculated for each analysis and was used as the main variable de-scribing the solid solution. The solid solution is then described forNi–Mg kerolite samples by the Ni percentage in the octahedralsheet, ranging from 0 to 100 (Ni0 to Ni100).

Transmission electron microscopy (TEM)

TEM photomicrographs, energy dispersive spectra (EDS), and elec-tron diffraction patterns were obtained on dried powder samplesdispersed in ethanol and deposited on a micro grid. The PhilipsCM20 electron microscope operated at 200 kV and was equippedwith an ultrathin window X-ray detector. Microchemical analysesof isolated clay particles were carried out to confirm at nanoscalethe chemical compositions previously determined by EMPA. Theanalyses (Si, Mg, Ni, Fe, Al, and Na) were carried out with an energydispersive X-ray analyzer (EDAX) attached to the electron micro-scope in nanoprobe mode over 40 s with a probe diameter of10nm. The KAB factors were determined using standards with amaximum error of 5% for each element. For each sample, the anal-yses were performed on isolated particles for which the layer spac-ing was measured by using HR-TEM images.

Raman microspectroscopy

Equipment

The Raman spectra were recorded using a LabRAM HR spectrome-ter (Horiba Jobin Yvon) equipped with a 1800g.mm�1 grating andan edge filter. The confocal hole aperture was of 500μm, the slit ap-erture of 50μm. The excitation beam was provided by an Ar+ laser(Stabilite 2017, Spectra Physics, Newport Corporation) at 514.53nmand a power of 200mW focused on the sample using a 50× objec-tive. The spot size was <1μm. Acquisition time, limited by a weakluminescence, ranged between 2 s and 6 s. The number of accumu-lations was set between 10 and 30 to optimize the signal-to-noiseratio within a reasonable acquisition time.

Classical signal decomposition

The decomposition of the Raman signal in the OH stretching vibra-tion region was carried out using the Peak Fitting Module of Origin7.5 software. Previously, a straight baseline was subtracted using

Figure 1. Example of texture andNi zonation in pimelite (BSE-SEM picture).

M. Cathelineau et al.

wileyonlinelibrary.com/journal/jrs Copyright © 2015 John Wiley & Sons, Ltd. J. Raman Spectrosc. (2015)

Page 128: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Labspec 5.64.15 software (Horiba Jobin Yvon). The number of peaks(9) was defined as the minimal number of peaks necessary to reachcoherent fits of the spectra on thewhole range of Ni content. As thepeaks were very close to each other, the peak shape was fixed to50–50 Gaussian–Lorentzian functions, and the variation of positionand full width at half maximum (FWHM) of each peak were also lim-ited in order to avoid large deviations of the decomposition overthe whole sample range. Peak areas are calculated from the fittingprocess and expressed as relative peak areas after normalization bythe sum of the peak areas of each spectrum.

Multivariate curve resolution

Following the chemometrics terminology, multivariate curve reso-lution (MCR) consists in decomposing a set of experimental spectracorresponding to the mixture of pure chemical species.[11] In thesignal processing community, it is also termed as non-negativesource separation.[12] Gathering the set of experimental spectra intoa matrix D of dimension (M,N), MCR consists in decomposing D ac-cording to:

D ¼ C ST þ E

where C and S are matrices of dimensions (M,K) and (N,K) respec-tively. The matrix S represents the spectra of the K pure chemicalspecies while matrix C describes how the contribution of the chem-ical species varies according to a physical parameter, also termed asdiversity. Here, diversity is the Ni substitution rate. E is the residualmatrix accounting for model errors. The decomposition of D wasperformed using an algorithm called Bayesian Positive SourceSeparation (BPSS)[12,13] which has proved to be very effective atseparating spectroscopic mixtures in difficult cases suffering fromrotational ambiguities (see Moussaoui et al.[13] for details). Prior to

the BPSS procedure, a baseline subtraction was performed[14] to re-move irrelevant information from the experimental spectra.

Results

Crystal chemistry of kerolites

Figure 1 provides an example of the zoned kerolite withcollomorph texture. Each layer is characterized by a specific Ni con-tent, the Ni content varying from 44 to 48% NiO (~90–96% of Nisubstitution rate) in this area. The crystal chemistry of the sampleswas calculated from EMPA and based on the chemical formula(Ni,Mg)3 + x(Si4� y)O10(OH)2·nH2O. Na, Ca, and K are present onlyin small amount and are neglected.[2] The set of samples coversthe entire range between Mg–kerolite and Ni–kerolite (pimelite),from 0 to 96% of substitution of Mg by Ni (Table 1). Chemical com-position was confirmed by TEM-EDS nanoscale analyses. The chem-ical compositions are constant on the whole series (except Ni/Mgsubstitution). Only the amount of Co seems to increase with theNi content, but the number of Co in the structural formula is gener-ally lower than 1%, with a maximum of 3% in the sample Ni96. Thetetrahedral site shows a deficit of occupancy with an average value(±2σ) equal to 3.72± 0.05 apfu (atoms per formula unit) (ideally of4), contrary to octahedral sites which average occupancy is equalto 3.54±0.11 apfu (ideally of 3). The average Si/(Mg+Ni + Si)atomic ratio is equal to 0.51± 0.01 and is constant whatever theNi content (Fig. 2 and Table 1). This atomic ratio is lower thanthe one of talc (0.57 in Mg3Si4O10(OH)2) and much higher thanthe one of the serpentine group minerals (0.40 in Mg3Si2O5(OH)4).It is also lower than the one of 0.56 calculated from the data ofBrindley et al.[2] In cases ofmixed layering between talc-like and ser-pentine, the Si/(Mg+Ni + Si) ratio varied between 0.5 and 0.4.[5,15,16]

Table 1. Chemical analysis of the set of samples calculated from EMPA

Sample number 1 2 3 4 5 6 7 8 9 10 11 12 13

Ni0a Ni0b Ni0c Ni15 Ni22a Ni22b Ni31 Ni47 Ni65 Ni71 Ni91a Ni91b Ni96

SiO2 54.32 47.45 57.89 47.74 49.36 47.97 48.24 43.23 46.39 41.05 39.53 41.40 45.58

Al2O3 0.00 0.02 0.00 0.08 0.00 0.43 0.07 0.04 0.01 0.04 0.00 0.48 0.00

MgO 35.29 36.68 33.04 25.87 22.99 23.99 20.12 15.58 10.14 7.28 2.58 2.35 1.06

K2O 0.00 0.00 0.00 0.00 0.06 0.02 0.06 0.03 0.00 0.01 0.00 0.02 0.01

Na2O 0.00 0.00 0.04 0.02 0.09 0.03 0.09 0.04 0.00 0.03 0.00 0.02 0.00

Cr2O3 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.03 0.00

MnO 0.00 0.00 0.00 0.00 0.00 0.00 0.06 0.01 0.01 0.00 0.00 0.10 0.02

FeO 0.02 0.00 0.02 0.00 0.29 0.12 0.39 0.14 0.06 0.29 0.06 0.01 0.00

CoO 0.00 0.08 0.04 0.00 0.09 0.09 0.00 0.03 0.02 0.26 0.44 1.23 1.50

NiO 0.00 0.05 0.06 8.44 12.03 12.41 16.84 25.28 34.96 33.61 46.33 45.01 48.51

Total 89.63 84.28 91.08 82.19 85.01 85.16 85.95 84.43 91.59 82.63 88.99 90.66 96.68

H2O

Si 3.71 3.49 3.86 3.72 3.79 3.70 3.77 3.65 3.75 3.75 3.60 3.66 3.77

Al 0.00 0.00 0.00 0.01 0.00 0.04 0.01 0.00 0.00 0.00 0.00 0.05 0.00PTetr. 3.71 3.49 3.86 3.73 3.79 3.74 3.78 3.65 3.75 3.75 3.60 3.71 3.77

Fe2+ 0.00 0.00 0.00 0.00 0.02 0.01 0.03 0.01 0.00 0.02 0.00 0.00 0.00

Co 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.00 0.00 0.02 0.03 0.09 0.10

Mg 3.59 4.02 3.28 3.01 2.63 2.76 2.34 1.96 1.22 0.99 0.35 0.31 0.13

Ni 0.00 0.00 0.00 0.53 0.74 0.77 1.06 1.72 2.27 2.47 3.40 3.20 3.23POct. 3.59 4.02 3.28 3.54 3.40 3.54 3.43 3.69 3.50 3.50 3.78 3.60 3.46

Ni/(Ni +Mg) % 0.00 0.00 0.00 15 22 22 31 47 65 71 91 91 96

Si/(Si + Ni +Mg) 0.51 0.46 0.54 0.51 0.53 0.51 0.53 0.50 0.52 0.52 0.49 0.51 0.53

OH stretching vibrations in Ni–Mg kerolite

J. Raman Spectrosc. (2015) Copyright © 2015 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs

Page 129: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Interlayer spacing

HR-TEM images indicate the presence of only one type of sheet sil-icate with 5 to 20 layers with similar interlayer distances of ~9.5Å(Fig. 3) with no evidence of mixed layering with 7-Å phases. Thespacing distance determined by electron diffraction is also around9.4± 0.1Å. Such values are similar to the one of the talc-like mineraldescribed in Falcondo deposits by Villanova-de-Benavent et al.[5]

which HR-TEM images provided an estimate of 9.5Å, and the onesprovided by Brindley et al.[2] with a value of 9.6Å. Basal spacing ofkerolite is thus slightly larger than that of talc (9.34Å) and may belinked to disordered layer stacking.[2] The absence of mixed layeredphases in the present samples is thus confirmed by crystal chemis-try and HR-TEM images.

Raman spectra of the Ni–Mg kerolite

The Mg–kerolite endmember

The three samples of Mg–kerolite endmember (Ni0) were analyzedin two regions, a low wavenumber (100–1200 cm�1), and a high

wavenumber region (3550–3750 cm�1) (Fig. 4). Below 1200 cm�1,the Raman spectrum of the Mg–kerolite Ni0 shows several bandsat similar position to the ones observed in both talc and chrysotile,except the one of chrysotile at 622 cm�1. Additional peaks are ob-served at 108, 187, and 840 cm�1. In the region of OH stretching vi-brations, the global shape of the band is more or less similar to theone of chrysotile, with a maximum at 3700 cm�1, but enlarged to-ward lower and higher wavenumbers. The peak at 3685 cm�1 re-ported in literature for chrysotile[17] is not visible in the Ramanspectrum of Mg–kerolite. No signal corresponding to water mole-cules (3000–3500 cm�1) or brucite layer (3654 cm�1[18]) is detected.

The Ni–Mg kerolite

Variations in the Raman spectra of these samples in the low wave-number region (100–1200 cm�1) and their structural implicationsare under way, but will be presented separately. The remainder ofthis study will focus on the OH stretching region. The Raman spec-tra in the OH stretching vibration region of seven selected samples

Figure 2. Si – Mg+ Fe – Ni ternary diagram (EMPA data).

Figure 3. HR-TEM picture of a talc-like mineral particle. Figure 4. Raman spectra of chrysotile, talc, and Mg–kerolite.

M. Cathelineau et al.

wileyonlinelibrary.com/journal/jrs Copyright © 2015 John Wiley & Sons, Ltd. J. Raman Spectrosc. (2015)

Page 130: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

are shown in Fig. 5. With increasing Ni content, the shape of the OHstretching vibration band shows a strong evolution, with a shift to-ward lower wavenumbers and an evolution to a complex bandstructure with several maxima. The position of the intensity maxi-mum of the band, first around 3700 cm�1, decreases, and threeother maxima successively appear and disappear at about 3685,3670, and then 3650 cm�1. In the middle of the substitution range(spectrum Ni47 in 5), the spectrum is a complex band cluster be-tween 3620 and 3700 cm�1 with three major peaks at 3649, 3670,and 3700 cm�1. For the highest Ni contents (Ni90 to Ni96), theRaman spectrum exhibits a thin strong peak at 3650 cm�1 with ashoulder at 3625 cm�1 and a larger one at 3660 cm�1.

Figure 6 shows some examples of spectrum decomposition intonine Gaussian–Lorentzian functions. The variations of the relativeareas of the nine peaks composing the Raman signal between3600 and 3750 cm�1 with Ni substitution rate are shown in Fig. 7.The areas of the 5 peaks between 3658 cm�1 and 3625 cm�1 in-crease with the Ni content, whereas the areas of the 2 peaks at3700 cm�1 and 3685 cm�1 decrease. The variations of area of thetwo peaks in an intermediary range at 3675 cm�1 and 3669 cm�1

are not so clear. The relative area of the peak at 3669 cm�1 startsat zero and increases up to a 60–70% of Ni substitution then possi-bly reaches a plateau. It possibly decreases at the very last steps. Therelative area of the peak at 3675 cm�1 seems to increase in the firststeps and then regularly decreases down to zero at the highest Nicontents. However, the strong superimposition of the nine peaksto each other and the constraints necessary to get consistent dataprocessing (fixed shape factor, limited variations in positions andFWHM) limit the interpretation of these variations.

In order to avoid over- ormis-interpretation of the spectra, and tolearn more about kerolite structure, the set of Raman spectra is an-alyzed using the BPSS chemometrics method.[12] After backgroundsubtraction[14] and normalization, the corrected Raman spectrawere arranged from top to bottom in order of the Ni–Mg substitu-tion rate (SR) of the corresponding sample (Fig. 8A). The three firstspectra show a maximum at ~3730 cm�1. In spectra # 4 to 6, asecond weak maximum appears at ~3710 cm�1. In sample # 7 to10, the intensity of the first maximum at ~3730 cm�1 decreases,and a third and fourth maxima are visible at ~3700 cm�1 and3675 cm�1. In the three last samples (# 11 to 13), the fourth maxi-mum at 3675 cm�1 becomes predominant. After data processing,a similar map is drawn with the spectra calculated from the model(Fig. 8B). The difference between the two maps results in the error

map (Fig. 8C). Errors are randomly distributed, symmetric to zero, andthe error intensitymaximum (0.02) of about 10%of the spectra inten-sity maximum (0.18). These results validate the calculated model.

The data of the calculated model are shown in Fig. 9A. By trialand error method, four coherent, interpretable sources are deter-mined to compose the Raman signal. The contribution of each ofthese sources varies with SR. Source #1 is predominant at 0% ofNi and regularly decreases to be at zero at about 65% SR. Source#2 first increases to reach a maximum at ~20% SR and then regu-larly decreases down to zero at ~70% SR. Source #3 regularly in-creases from zero to reach a maximum of contribution in thesignal at ~70% SR and then decreases down to zero at 90% SR.Finally, source #4 is equal to zero up to 20% SR and then regularlyincreases to be predominant at 90% SR. The Raman spectra of thefour sources are shown in Fig. 9B. The peak maxima moves fromhigh wavenumbers to lower wavenumbers with increasing Ni con-tent, in agreement with the higher mass of Ni than Mg. The shapeof the bands is complex, with several local maxima except in source#1. Source #1 and source #2 maxima overlapped but source #2 pre-sents three other peaks.

Discussion

Structural formula of the Ni–Mg kerolite

The crystallographic structure of kerolite has been proposed to besimilar to the one of talc, with a TOT sheet arrangement (2:1), the

Figure 5. Normalized Raman spectra of some Ni–Mg talc-like mineralsfrom 0 to 96% Ni substitution rate in the region of OH stretching vibrations.

Figure 6. Decomposition into Gaussian–Lorentzian functions of the Ramanspectra of four Ni–Mg talc-like mineral samples.

OH stretching vibrations in Ni–Mg kerolite

J. Raman Spectrosc. (2015) Copyright © 2015 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs

Page 131: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

interlayer size being of ~9.5 Å.[2] Ni progressively substitutes to Mgin the octahedral sites without changing the interlayer spacing.Ni–Mg kerolite–pimelite series can thus be presented as a solid so-lution. However, the chemical analysis shows some deviation froman ideal chemical formula, with a deficit in tetrahedral site occu-pancy, an excess of octahedral site occupancy, and a total oxideweight percent lower than 100% (Table 1). The defect rate andthe total oxide weight percent are not correlated to the Ni content.The deviation from ideal formula was already reported in previousstudies.[1,9] Wells et al.[1] did not provide any explanation for this de-viation, excluding phase mixing from XRD data and homogeneoustexture at the observation scale (SEM) as in the present study.The chemical formula of pure Mg–kerolite, Mg3Si4O10(OH)2·nH2O

suggested by Brindley et al.[2] is the same as the one of talcMg3Si4O10(OH)2, except water. However, the Raman spectrum ofMg–kerolite is far from the one of talc (Fig. 4) and no water is ob-served. In talc one sixth of the oxygen atoms is protonated andbonded to one H and three Mg atoms. The thin peak at3675 cm�1 is linked to the stretching vibration of the only one typeof hydroxyl. In Mg–kerolite, the Raman spectrum is very different tothe one of talc, even in pure Mg–kerolite. It is indicative of a differ-ent crystal chemical formulae. The deviation from ideal chemicalformula, based on the one of talc, and Raman spectroscopy thusboth show that the structure of the two minerals is somewhat dif-ferent. This difference might be related to the perturbation of thehydroxyl groups by Ni–Mg arrangement, stacking defects, or pres-ence of bound water in the structure (interlayer, surface).[8] TheRaman spectra are also quite distinct from those of serpentine,thereby allowing rejecting the hypothesis of mixed layering ormixing with serpentine.

Raman spectra of OH stretching vibrations in Ni–Mg kerolite

The Raman spectra of Ni–Mg kerolite in the OH stretching vibrationregion are extremely complex. If four maxima are visible, no individ-ual peaks clearly appear in the band structure. The band shapestrongly evolves with the chemistry of the mineral. Chemometricsapproach clearly distinguished four sources. Following the coher-ent and continuous evolution of the contribution of these sourceswith Ni content, each of them can be assigned to one environmentof the OH groups, i.e. one arrangement of the Ni and Mg cations inthe octahedral sites: (1) Mg3; (2) Mg2Ni; (3) MgNi2; and (4) Ni3, theposition of the OH bands shifting to lower wavenumbers whenthe Ni content increases. However, the second source correspond-ing to theMg2Ni arrangementwhich appears with the first Ni atomscannot be fully differentiated from source #1 as the peak maximumof the two sources coincides. It might be because of a lack of sam-ples in the range 1–8% NiO (1–15% SR) where the contribution ofsource #2 might range between 5 and 30%. Ideally the set ofsamples would be completed by samples of Ni contents in therange 1–15% SR to better distinguish between source #1 and #2.Unfortunately, the only samples we found in this range presenteda different crystal chemistry, mainly a different Si/(Si +Mg+Ni) ratio(Fig. 2) and different Raman spectra, and thus were not included inthe present study.

The sources are simultaneously present in the spectrum over awide range of Ni content. For example, at 50% SR, each of the foursources contributes for about 25% of the Raman signal. Conse-quently, the distribution of Ni andMg in the octahedral site is prob-ably heterogeneous, with areas still free of Ni at relatively high Nicontent, and Ni-rich areas in the first steps of Ni–Mg substitution.

Figure 7. Evolution of the relative area of the Raman peaks in the OH stretching vibration region as a function of Ni–Mg substitution rate.

M. Cathelineau et al.

wileyonlinelibrary.com/journal/jrs Copyright © 2015 John Wiley & Sons, Ltd. J. Raman Spectrosc. (2015)

Page 132: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

It validates one of the hypotheses stated in previous studies[3,9] onthe basis of EXAFS and IR studies of kerolite–pimelite series.Manceau and Calas[3] suggested the possibility of mixed layeringbetween Mg– and Ni–kerolite layers or the clustering of Ni in theoctahedral sheet. From the present result, mixed layer hypothesiscan be rejected. The presence of Ni clusters perfectly fits with thepresent Raman spectra, as suggested by the IR spectra of Gerardand Herbillon.[9]

In an infrared study of synthetic talcs of various Ni contents,Wilkins and Ito[10] showed four peaks in the OH stretching vibrationrange, directly linked to the four possible arrangements in theoctahedral layer. Here the shape of the Raman signal of the OHstretching vibration does not consist of single well-defined peaksbut presents a complex bands structure. It reveals an OH environ-ment more complex that the one of talc (taken as model structure),even in pure Mg kerolite. The complexity of the signal might be

0 10 20 30 40 50 60 70 80 90 1000

0.5

1

1.5

3640 3660 3680 3700 3720 3740 3760 37800

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16Mixing A Sources S

(a) (b)

Figure 9. Variations (A) of the contributions to the Raman signal of the four sources (B) as a function of Ni substitution rate Ni/(Ni +Mg) (%).

Mixtures

3660 3680 3700 3720 3740 3760

2

4

6

8

10

120

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18Model

3660 3680 3700 3720 3740 3760

2

4

6

8

10

12 0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

Error

3660 3680 3700 3720 3740 3760

2

4

6

8

10

12 −0.015

−0.01

−0.005

0

0.005

0.01

0.015

0.02

(a)

(c)

(b)

Figure 8. Startingmixturemap (A), result map (B), and error map (C) from BPSS data processing. X-axis: wavenumbers/cm�1; y-axis: sample number (cf. Table 1).

OH stretching vibrations in Ni–Mg kerolite

J. Raman Spectrosc. (2015) Copyright © 2015 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs

Page 133: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

explained by the appearance of more than one OH site symmetry(contrary to talc) because of structural defects in the octahedraland tetrahedral sheets, stacking disorder, vibrational couplingbetween hydroxyls, and/or surface OH/H2O molecules. Betterunderstanding of OH arrangement might be provided by recordingOD Raman spectra of kerolite after H-D exchange by exposingsamples to D2O vapor, to dissociate potentially coupled OH groupsor surface groups. However, obtaining sufficiently low deuteriumconcentration by this method is not trivial.[19,20]

Conclusion

A set of Raman spectra of Ni–Mg kerolite over the complete Ni–Mgsubstitution range was recorded. The present study shows Ni–Mgkerolite to occur as a specific individual mineral phase with a con-stant crystal chemistry over the whole Ni–Mg composition rangeand with a constant interlayer spacing, distinct from talc andserpentine. Mixed layering with other phases (brucite, serpentine)is excluded to explain the shift from ideal stoichiometry. Althoughkerolite and pimelite have not been formally recognized as distinctminerals by the IMA this study shows that these phases exist asdistinct minerals and not as mixed layering/mixed mineral mate-rials. Thanks to the combination of Raman spectroscopy withchemometrics, chemical analysis, and HR-TEM observations, somestructural features of kerolite can be determined. Ni andMg cationsare in the octahedral sheet and are heterogeneously distributed,forming clusters of Mg and Ni even at low Mg or Ni content. Fourarrangements of Ni andMg in octahedral sites are identified and re-lated to a specific Raman signal. The Raman signals related to eachgeometry are not single peaks but broad complex band structure,highlighting a more complex OH structure than the one suggestedby homology with talc, in agreement with the deviation from idealformula calculated from chemical analysis.Raman spectroscopy reveals to be an excellent tool for

deciphering the complex Ni–Mgphyllosilicate association in garnier-ite Ni-ores, when used in close association with other macroscopicto microscopic observations, and when signal is not hidden by fluo-rescence. The use of in-situ techniques at micro-scale such as Ramanspectroscopy and electron microprobe analyses makes possible toovercome the problem of mixing encountered by XRD and IR orNIR spectroscopy. Raman spectroscopy is thus a simple techniqueto clearly distinguish between talc, kerolite, and serpentine. The highsensitivity of OH groups Raman signature to the Ni content offersthe possibility to use this technique in the field for quick recognitionof local Ni enrichment, which should open new perspectives withthe development of portable micro-Raman techniques.

Acknowledgements

This work has been carried out thanks to the financial help of LabexRessources 21 (supported by the French National Research Agency

through the national program ‘Investissements d’avenir’ with refer-ence ANR-10-LABX-21—LABEX RESSOURCES 21), and with the ana-lytical capabilities of the GeoRessources laboratory (Université deLorraine, CNRS, CREGU). A few samples were first collected duringan early introduction on the first one day visit at Koniambo andPoro in 2011 led by C. Couteau and E. Fritsch. The field trip was or-ganized under the framework of the CNRT (French National Centrefor Technological Research) project ‘Nickel and its environment’. Asecond series of samples were collected during two-week stays in2012 and 2013 financed by Koniambo within the framework ofthe PhD of B. Quesnel, and supervised in the field by C. Couteauand M. Drouillet from Koniambo S.A. This field works includedtwo visits to Poro mine of SLN which is also warmly acknowledged.The authors acknowledge J. Ghanbaja for the acquisition of HR-TEMimages and S. Mathieu and O. Rouer for their help in the character-ization by SEM. The authors are also grateful to two anonymousjournal reviewers for their helpful comments and to Dr. CraigMarshall for editorial work.

References[1] M. A. Wells, E. R. Ramanaidou, M. Verrall, C. Tessarolo, Eur. J. Mineral.

2009, 21, 467.[2] G. W. Brindley, D. L. Bish, H.-M. Wan, Am. Mineral. 1979, 64, 615.[3] A. Manceau, G. Calas, Am. Mineral. 1985, 70, 549.[4] A. Manceau, G. Calas, Clay Miner. 1986, 21, 341.[5] C. Villanova-de-Benavent, J. A. Proenza, S. Galí, A. Garcia-Casco,

E. Tauler, J. F. Lewis, F. Longo, Ore Geol. Rev. 2014, 58, 91.[6] C. Villanova-de-Benavent, T. Aiglsperger, T. Jawhari, J. A. Proenza,

S. Galí, in macla n°16, (Eds.: J. M. Herrero, J. Arostegi, P. P. Gil),Congreso SEM-SEA, Sociedad Española De Mineralogía, Bilbao, 2012,pp. 180–181. ISSN 1885-7264

[7] M. Cathelineau, M. Harlaux, R. Mosser-Ruck, M.-C. Caumon, P. Boulvais,E. Fritsch, C. Couteau, GeoRaman 10, Nancy, France, 2012, pp. 111–112.

[8] G. W. Brindley, Mineral. Mag. 1977, 41, 443.[9] P. Gerard, A. J. Herbillon, Clays Clay Miner. 1983, 31, 143.

[10] R. W. T. Wilkins, J. Ito, Am. Mineral. 1967, 52, 1649.[11] A. de Juan, R. Tauler, Crit. Rev. Anal. Chem. 2006, 36, 163.[12] S. Moussaoui, D. Brie, A. Mohammad-Djafari, C. Carteret, IEEE Trans.

Signal Process. 2006, 54, 4133.[13] S. Moussaoui, C. Carteret, D. Brie, A. Mohammad-Djafari, Chemometr.

Intell. Lab. Syst 2006, 81, 137.[14] V. Mazet, C. Carteret, D. Brie, J. Idier, B. Humbert, Chemometr. Intell. Lab.

Syst. 2005, 76, 121.[15] G. Dublet, F. Juillot, G. Morin, E. Fritsch, D. Fandeur, G. Ona-Nguema,

G. E. Brown Jr., Geochim. Cosmochim. Acta 2012, 95, 119.[16] G. Dublet, Relation entre spéciation et distribution du Nickel dans les

couvertures d’altération latéritique des roches ultrabasiques deNouvelle-Calédonie, PhD thesis, Paris 6, 2013.

[17] D. Bard, J. Yarwood, B. Tylee, J. Raman Spectrosc. 1997, 28, 803.[18] B. Weckler, H. D. Lutz, Spectrochim. Acta A Mol. Biomol. Spectrosc. 1996,

52, 1507.[19] M. C. Jodin-Caumon, B. Humbert, N. Phambu, F. Gaboriaud,

Spectrochim. Acta A Mol. Biomol. Spectrosc. 2009, 72, 959.[20] H. D. Lutz, J. Mol. Struct. 2003, 646, 227.

M. Cathelineau et al.

wileyonlinelibrary.com/journal/jrs Copyright © 2015 John Wiley & Sons, Ltd. J. Raman Spectrosc. (2015)

Page 134: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 135: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genesis and mineralogy of Ni ores

Cathelineau, M., Quesnel, B., Gautier, P., Boulvais, P., Couteau, C., Drouillet, M., (2016b).

Nickel dispersion and enrichment at the bottom of the regolith: formation of pimelite target-like ores in rock block joints

(Koniambo Ni deposit, New Caledonia)

Mineralium Deposita 51, 271–282. doi:10.1007/s00126-015-0607-y

- 10 -

Page 136: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

ARTICLE

Nickel dispersion and enrichment at the bottom of the regolith:formation of pimelite target-like ores in rock block joints(Koniambo Ni deposit, New Caledonia)

Michel Cathelineau1& Benoît Quesnel2 & Pierre Gautier2 & Philippe Boulvais2 &

Clément Couteau3& Maxime Drouillet3

Received: 5 November 2014 /Accepted: 23 July 2015 /Published online: 9 August 2015# Springer-Verlag Berlin Heidelberg 2015

Abstract In New Caledonian Ni deposits, the richest Ni sili-cate ores occur in fractures within the bedrock and saprolite,generally several tens of meters to hundred meters below thepresent-day surface. Fracture-related Ni silicate ore accountsfor high Ni grades, at least a few weight percent above theaverage exploited grade (2.5 %). These Ni-rich veins are af-fected by active dissolution-precipitation processes at the levelof the water table. Ni in solution is precipitated as silicates inthin layer cementing joints. This mineralization is character-ized by chemical and mineralogical concentric zoning with anouter green rim around an inner white zone composed, fromthe edge to the centre of the block, (i) a highly oxidized andaltered zone, (ii) a green pure Ni-rich pimelite zone, (iii) azone (limited to a few centimetres) with a mixture of Ni-poor kerolite and Ni-rich pimelite and intermediate coloursand (iv) a large white Mg-kerolite mineralization zone. Thisstudy proposes that the concentric zonation results fromevapo-precipitation process related to alternate periods of hy-dration and drying, induced by water table movements. Thisextensive dispersion of Ni in concentrically zoned ores canpartly explain the rather monotonous Ni grade of the bulk

exploitation at the base of the regolith with values between 2and 3 wt%.

Keywords Kerolite-pimelite . Garnierite . Ni-laterite . NewCaledonia . Dissolution-precipitation

Introduction

Nickel silicate ores are generally thought to be closely associ-ated with the weathering of peridotite leading to the formationof laterite profile (Trescases 1975; Schellman 1983; Gleesonet al. 2003; Freyssinet et al. 2005; Villanova-de-Benaventet al. 2014). In New Caledonia, the main ore genetic modelfor Ni ores is based on a single per descensum model wheremost elements (Mg, Ni and Si) are leached from the surface(Trescases 1975; Troly et al. 1979; Pelletier 1983, 1996).Nickel is then concentrated either in the fine-grained lateritewhere goethite is the main Ni bearer, the so-called lateritic ore(Manceau et al. 2000; Dublet et al. 2012), or below the later-ites sensu stricto in the regolith as mixtures of hydrous Mg-Nisilicate and goethite, the so-called saprolite ores (Dublet et al.2012). Ni-rich silicates also occur as fracture fillings, some-times considered as syn-tectonic (Cluzel and Vigier 2008).These veins are filled with Bgarnierite^, which is a term usedfor a mixture of different Ni silicates: mainly the kerolite-pimelite series, also called talc-like, as well as sepiolite andNi-serpentine, also called Bserpentine-like^ (Brindley andHang 1973; Springer 1974; Brindley and Maksimovic1974). In most studies, Ni-rich silicates were described indetail using a series of appropriate analytical tools (X-ray dif-fraction (XRD) and infrared spectroscopy (IR) in some casesin combination with microprobe studies (Brindley and Hang1973; Brindley and Maksimovic 1974; Brindley and Wan1975; Brindley et al. 1977, 1979; Wells et al. 2009; IR:

Editorial handling: T. Bissig and G. Beaudoin

* Michel [email protected]

* Benoît [email protected]

1 Université de Lorraine, GeoRessources Laboratory, CNRS, CREGU,BP 70239, F-54506 Vandoeuvre-lès-Nancy, France

2 Géosciences Rennes, Université Rennes 1, UMR 6118 CNRS,35042 Rennes Cedex, France

3 Service géologique, Koniambo Nickel SAS, 98883 Voh, NouvelleCalédonie, France

Miner Deposita (2016) 51:271–282DOI 10.1007/s00126-015-0607-y

Page 137: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Gerard and Herbillon 1983), extended X-ray absorption finestructure (EXAFS; Manceau and Calas 1985, 1986; Manceauet al. 1985) and some pioneer Raman works (Villanova- de-Benavent et al. 2012). Very few recent studies of Caledonianores are available, apart those fromWells et al. (2009), Dublet(2012) and Cathelineau et al. (2015). Generally speaking, thegeometric, time-space and paragenetic description of the oresis lacking.

Ni silicate ores have not been studied in detail at theKoniambo mining site (Fig. 1), which was recently re-opened as part of the Koniambo Nickel project, with the ex-ception of the descriptions provided by Fandeur (2009).Recent observations in the new open pits have shown thattwo main types of Ni silicate occur. The first type (type 1)consists of mineralized faults and fractures filled by Ni sili-cates a few centimetres to decimetres in width, typically cor-responding to what is described elsewhere in New Caledonia.These veins are filled by Ni-Mg kerolite-pimelite solid solu-tion (as defined by Brindley et al. 1977). These minerals weredescribed by Wells et al. (2009) at Goro (southern NewCaledonia, Fig. 1) and Dublet (2012) at Koniambo. The Ni-Mg kerolite-pimelite series is characterized by a lightgreenish-blue colour. They correspond to the so-called garni-erite fracture infilling considered as syntectonic by Cluzel andVigier (2008). In the case of Koniambo, the term garnieriteseems inappropriate, as the fracture ores are not mixtures ofseveral Ni-rich silicates such as serpentine, kerolite and sepi-olite as expected for garnierite but instead are only filled bybluish kerolite. The second type (type 2), which has not been

described so far, consists of chemical and mineralogical con-centric zoning, similar to a shooting target, and characterizedby an outer green rim around an inner white zone coatings onjoint planes that occur in the upper part of the open pits. Theconcentric zonation consists, from the edge to the centre of thejoint, (i) a highly oxidized and altered zone, (ii) a green zone,(iii) a zone (limited to a few centimetres) with an intermediatecolours and (iv) a large white zone. As there are now excep-tional outcrops in the Koniambo Nickel mining area wheretwo Ni ore types can be observed, a detailed structural andmineralogical study of the main ores has been carried out,especially on the so-called Cagou pit within the Koniambomining domain (Fig. 1). The main objective of this study isto understand the genetic process of target-like ore formation.We determined the detailed mineralogy of the target-like oresand documented the geometric and time-space relationshipsbetween the two ore types at the field scale. We provide a newconceptual model of the Ni distribution, which takes into ac-count the contribution of the two types of Ni silicates to thebulk grade of the exploited ore at the base of the regolith.

Geological setting

New Caledonia, in the SW Pacific, is an island characterizedby the presence of significant relicts from the obducted peri-dotite nappe overlying basement formations from the NorfolkRidge micro-continent (Fig. 1). Today, the Peridotite Nappe isessentially exposed in theMassif du Sud covering much of the

Fig. 1 a Simplified geological map of New Caledonia modified afterCluzel et al. (2001). The laterites are compiled from Paris (1981). bGeological map of the Koniambo massif modified after Maurizot et al.

(2002). The yellow diamond indicates the location of the study site(Cagou pit from the Koniambo Nickel mining company) in theKoniambo massif

272 Miner Deposita (2016) 51:271–282

Page 138: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

south-eastern third of the Grande Terre and as klippes alongthe north-western coast, including the Koniambo massif(Fig. 1). The south-westward emplacement of the PeridotiteNappe is considered to have occurred between 37 and 27 Ma(Cluzel et al. 2001, 2012; Paquette and Cluzel 2007; Maurizotet al. 2009). Outcropping peridotites then quickly underwentstrong water-rock interactions under a hot and humid climate,favouring the formation of a thick laterite cover (Fig. 1;Trescases 1975). Laterite formation consists of the progressivedissolution of all silicates found in dunite and harzburgite (e.g.olivine and orthopyroxene) and the development of residuallaterite, containing insoluble elements such as trivalent ironre-precipitated as goethite and/or hematite, and relicts of mag-matic chromite. The accumulation of goethite over a signifi-cant period of time produced a 5 to 50 m deep profile includ-ing a lateritic duricrust and a plasmic horizon over a ferrugi-nous saprolite, following the terminology from Butt andCluzel (2013). Palaeomagnetic ages (around 25 Ma) obtainedon lateritic duricrust from the Thiebaghi area (location inFig. 1) suggest that the main laterite formation took placerather early, from the late Oligocene to the Miocene (Sevinet al. 2012).

Presently, the hydrologic and topographic features of theoutcropping peridotite nappe are likely different to those of themain laterization stages. Mechanical erosion, helped by theconstant hydrolysis of the peridotite minerals, predominatesin topographically high areas, favoured by high orographicrainfall, with precipitation ranging from 2500 to 4000 mm(Printemps et al. 2007; Terry and Wotling 2011).

The Koniambo massif has been exploited intermittentlysince the end of the nineteenth century and particularly atthe end of the twentieth century by SMSP. In 1998, theKoniambo Nickel project was started as part of a SMSP-Falconbridge venture and was followed by the industrial pro-ject conducted by SMSP-XSTRATA Nickel after 2007.Koniambo hosts around 158.6 million tonnes of saprolite thatgrades 2.47 wt%Ni at a cut-off grade of 2 wt% (Xstrata 2012).

The Koniambo massif is characterized by hills culminatingat 930 m and steep slopes. At Koniambo, as in other parts ofNew Caledonia, the distribution of alteration facies does notfollow the simple flat geometric distribution of the lateriticprofile. The saprocks are overlain by oxidized horizons, most-ly ferruginous saprolite. Above the ferruginous saprolite, ontopographic highs, around 850 m above sea level, relicts ofweathering profiles are preserved. The remnants of the lateriticprofile, such as lateritic duricrust or pisolitic horizons as wellas the plasmic horizon, are discontinuous and are in partredistributed due to slope effects following the degradationprocess affecting the laterite profile described elsewhere inNew Caledonia (Chevillotte et al. 2006). The recent topresent-day conditions are characterized by active dissolutionprocesses which are remarkably illustrated by dissolutionpipes within hard serpentinized peridotite. Highly

transmissive karstic networks occur at depth. The local col-lapse is marked by sinkholes in the topography, as alreadyobserved elsewhere in New Caledonia (Trescases 1975;Genna et al. 2005). To summarize, two distinct conditionsexplain the geometric distribution of the alteration facies: (i)the present-day conditions, which mostly results in the disso-lution and oxidation of the serpentinized ultrabasic rocks in adynamic system which excludes the formation of a lateriteprofile due to permanent erosion, and (ii) older (Cenozoic)conditions which produced the lateritic profile. These profilesare almost entirely disturbed due to the competing effects oferosion and uplift and the subsequent flowing and reworking.

The studied zone, the Cagou pit (Fig. 1), is an open pitwhich provides a good cross-section of the lateritic profileincluding the bedrock, saprock and lower part of the ferrugi-nous saprolite. As in all of the upper part of the Koniambomassif, the bedrock and saprocks are highly fractured. Somefractures and joints in the saprock contain Ni-Mg hydroussilicates (Fig. 2a, b). The ferruginous saprolite above thesaprock contains significant Ni concentrations, which are re-lated both to oxihydroxides (Ni-goethite) and a fine-grainedNi silicate, inferred as a Ni-rich serpentine (Dublet et al.2012). These authors indicated that it is almost impossible toprecisely identify this Ni silicate phase, even using EXAFS.

Materials and methods

Field studies and sampling

A structural description and measurements of most joints andfractures were carried out in the Cagou pit based on the sys-tematic profiles along the approximately E-W pit face.Representative samples were taken in fractures, as well as onjoints presenting the target-like ores, some of which are shownin Fig. 2. The target-like ores consist of thin layers, generallyless than 500 μm thick. Therefore, most of the analyticalworks were carried out on sufficiently thick layers preparedas polished sections, with the exception of the Raman analysiswhich can be performed, without any preparation, on mostsamples.

Analytical techniques

The studied minerals are very fragile, and therefore, it wasdifficult to make polished thin sections. The layers were thusembedded in resin to establish the mineral paragenesis. Weused a conventional reflected light microscope and scanningelectron microscope (SEM), a JEOL equipped with an energydispersive spectrometer using a Si(Li) semi-conductor detec-tor and a HITACHI S-4800 SEM at SCMEM (Nancy, France).Minerals were identified based on the combined considerationof in situ chemical data using both semi-quantitative energy

Miner Deposita (2016) 51:271–282 273

Page 139: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

dispersive (EDS) data using SEM and a quantitative analysisusing an electron microprobe (EMPA), transmission electronmicroscopy data (TEM), and Raman spectroscopy data.

Electron microprobe analyses (EMPA) of the kerolite wereperformed at SCMEM (Nancy, France). Si, Al, Fe, Mn, Mg,Co, Ni and Na were analysed using a CAMECA SX100 in-strument equipped with a WDS and calibrated using naturaland synthetic minerals or compounds such as albite (Si, Na),

Al2O3 (Al), olivine (Mg), hematite (Fe), MnTiO3 (Mn), Co(Co) and NiO (Ni). Only traces of Al, Mn, Co and Fe werefound. The analytical conditions were a current of 12 nA, anaccelerating voltage of 15 kVand a counting time of 10 s (and30 or 60 s for Ni in Ni-poor kerolite). The analyses have aspatial resolution of 1 to 2 μm. The total Fe is presented asFeO. Structural formulae of kerolite-pimelite were calculatedarbitrarily on the basis of 22 negative charges per half unit cell,

Fig. 2 a Sketch summarizing field observations on the studied cross-section in the Cagou pit. The green, yellow and orange coloursrepresent the zones variously affected by dissolution and oxidizingprocesses. The location of the type 1 and type 2 ores is also reported. bField view of the studied cross-section in the Cagou pit (Fig. 1). c Equal-

area lower hemisphere stereographic projection of the kerolite veins andtarget-like bearing joints at the open pit scale. d Whole rock majorelement distribution along a drill hole located along the dashed line.Each chemical composition is representative of a given thickness ofrock represented on the sketch by a double black arrow

274 Miner Deposita (2016) 51:271–282

Page 140: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

i.e., an O10(OH)2 base, because the TEM observations andXRD data have not shown any other mineral layer than thekerolite-pimelite series. IMA’s 50-50 rule for solid solutionwas applied, e.g., kerolite was used from a Ni/Ni + Mg ratioof 0 to 0.5, and pimelite from a ratio of 0.5 to 1, although thesolid solution, in particular for ore type 1 consists of nearlycontinuous Ni/Ni + Mg ratios in between 0.2 and 0.7. Thetetrahedral sheet was assumed to be filled with Si. The octa-hedral sheet was filled with Ni and Mg.

TEM photomicrographs, energy dispersive spectra andelectron diffraction patterns were obtained on dried powdersamples dispersed in ethanol and deposited on a micro grid(Formvar/Carbon 300 Mesh Ni, Agar Scientific, Essex,England). The Philips CM20 TEM operated at 200 kV andwas equipped with an ultrathin window X-ray detector.

Raman spectra were recorded using a LabRAM HR spec-trometer (Horiba Jobin Yvon) equipped with an 1800 g mm−1

grating and an edge filter. The confocal hole aperture was500 μm, and the slit aperture was 50 μm. The excitation beamwas provided by a Stabilite 2017 Ar+ laser (Spectra Physics,Newport Corporation) at 514.53 nm and a power of 200 mW,focused on the sample using a ×50 objective (Olympus). Thespot size was less than 1 μm. The acquisition time was limitedby a weak luminescence, ranging between 2 and 6 s. Thenumber of accumulations was set between 10 and 30 in orderto optimize the signal-to-noise ratio (S/N).

Results

Mode of occurrence of the ores in the field

Figure 2b shows an E-W oriented profile in the Cagou pit,which shows the dissolution-oxidation front (reddish-brownish colours) and the underlying bedrock (whitish-palebrown serpentinized basic rocks). The overall saprock is af-fected by extreme fracturation, which is dominated by openednetworks of subvertical discontinuities, connected with lowangle joints. The main alteration zones and ore types are de-scribed below from the bottom of the regolith to the bedrock,based on Fig. 2a, b.

In the yellow zone, highly microfissured rocks have under-gone significant rock dissolution. This is where hydrolysis ofthe main silicates that constitute the peridotite protolith (serpen-tine and relicts of olivine) takes place. The silica dissolved fromthe upper levels recrystallizes here as numerous quartz layersand microfissure infillings, explaining the high local SiO2 con-tents (50 to 60 wt%). In the orange zone, saprocks are impreg-nated by iron oxihydroxides among which Ni-goethite is thepredominant Ni-bearer, as already described in the Koniambosaprolitic horizons by Dublet et al. (2012). In the green zones,which are less affected by oxidation and dissolution, two typesof joints and fractures bearing Ni silicates are observed: (i)

fractures with kerolite-pimelite infillings (ore type 1) and (ii)target-like joints (ore type 2). A collapse breccia is also ob-served and corresponds to a pipe filled by centimetres todecimetres rock blocks that fell down into the pipe and werepartially cemented by quartz and Ni-rich pimelite.

Figure 2c is the stereographic projection of the dip anddirection of the discontinuities filled by the two main silicateore types. Type 1 discontinuities split into two families with astrike close to N150 with a dip ~85 E-W for the first and closeto N50 with a dip range between 30 N-Wand 80 N-W for thesecond. Type 2 ore-bearing joints occur in nearly cementedjoints without systematic strike and dip. They constituteplanes in which longest length ranges from 30 cm to 1 m,rarely more than 1 m, and occur in the vicinity of the type 1ore-bearing fractures. These target-like ore joints are revealedduring exploitation of the open pit when rock blocks are bro-ken by excavators. This observation possibly indicates thatthese mineralized joints represent closed and filled fractures.When opened during quarrying, target-like ores are quicklyaltered by present-day meteoric waters, hydrolysis inducingrapid mechanical and chemical degradation.

Figure 2d provides the NiO concentrations along a drillcore, located on the studied profile (Fig. 2a) and taken beforethe pit was exploited. The NiO content ranges from 2.7 to5.4 wt% in the main zone of the target-like occurrences.

Textural description of the ores

Type 1 ores consist of centimetres to decimetres thickfracture infillings containing mostly kerolite-pimelite sol-id solution (Fig. 3a). It may be affected later on by thereopening of the vein and sealing by red to brownishmicrocrystalline quartz and clear quartz microfractures.The texture of the kerolite-pimelite filling consists of suc-cessive botryoidal growth bands of bluish-green coloursimilar to those described in Villanova-de-Benavent et al.(2014) for type IV garnierite in Dominican Republic ores.Type 2 target-like Ni ores occur exclusively as a layer inthin joints. Their extension ranges from a few decimetres to1 m (Fig. 3c, d), corresponding to the average size of therock blocks occurring within the first hundred meters be-low the surface (Fig. 2b). In the joint plane, pimelite ispresent as a green rim, whereas the centre of the joint isfilled by whitish, translucent Ni-free kerolite (Fig. 3c, d).The main precipitation zone for the Ni-rich pimelite islimited to the first ~15 to ~20 cm at the edge of the rockblock (Fig. 3c, d). Target-like features are concentric with,from the edge (close to the fracture) to the centre of theblock, (i) a highly oxidized and altered zone (Fig. 3c, e), (ii)a pure Ni-rich pimelite zone (Fig. 3c–e), (iii) a zone (lim-ited to a few centimetres) with an apparent mixture of Ni-poor kerolite and Ni-rich pimelite and intermediate colours(Fig. 3c–e), (iv) a large Mg-kerolite mineralization zone

Miner Deposita (2016) 51:271–282 275

Page 141: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

(Fig. 3c–e) and (v) a poorly filled zone with a very thinMg-kerolite layer (Fig. 3c, e). In zone iii, the intermediatecolours result in fact from the superimposed deposit of thin

layers of green pimelite and Ni-poor white kerolite. All thekerolite infillings form successive thin layers on the rockand have a botryoidal texture (Fig. 3g, h).

Fig. 3 Field occurrence of type 1and 2 ores in the Cagou pit (seeFig. 1). a Type 1 kerolite-hematitebearing microcrystalline quartz(type 1 ore). b A relict of the type1 kerolite-hematite bearingmicrocrystalline quartz partiallydissolved by supergene alteration.c, d Zones of intense dissolutionand the remaining harzburgiteblocks showing, when brokenduring exploitation, a joint with atarget-like (ore type 2). Thecolour green corresponds to theNi-kerolite (pimelite) and thecolour white to the Mg-kerolite. eSketch summarizing themineralogical variation from theedge to the centre of the target-like. f A relict of the target-like(ore type 2) quickly altered afterexposure. g, h Zoom on the Mg-kerolite (g) and Ni-kerolite(pimelite) zones of a target-likeore showing the botryoidalhabitus

276 Miner Deposita (2016) 51:271–282

Page 142: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Both ore types are at present day affected bymeteoric waterpercolation. The water circulating into ore type 1 fracturesdissolves the Ni-Mg kerolite and partially dissolves the redto brownish microcrystalline quartz. Figure 3b shows the re-sult of the supergene alteration of the type 1 ores: The relictconsists of a totally unstructured vein, with reddish walls dueto goethite and hematite impregnations. Furthermore, Fig. 3fshows that target-like ores are also affected by the present-dayhydrolysis bymeteoric water quickly after exposure to surfaceconditions. The relict consists of a partially dissolved andmechanically dismembered target-like ore.

Mineralogy and crystal chemistry

All the studied minerals (Table 1, Figs. 3, 4, and 5) arekerolite, also called talc-like, as defined by Brindley andWan (1975) and Brindley et al. (1977). They are all character-ized by a broad peak on the XRD diffractograms around 10±0.5 Å due to poor crystallinity and small coherent domains; asusual, this value is slightly shifted from the interlayer valuethat was determined using HRTEM (9.5 Å). The type 1 oredoes not show any XRD peak corresponding to serpentine,but the occurrence of a peak between 7.26 and 7.30 Å is

frequently present in the type 2 ores and attributed to smallsized inclusions of serpentinized host rocks within the thinjoint infillings, which are impossible to separate from thekerolite. Raman spectra of the Ni-rich phases (pimelite fromore types 1 and 2) are characterized by a complex OH band inthe high frequency range (Fig. 4). This band is totally distinctfrom the spectra of all serpentine polymorphs. Raman spectraof Mg-kerolite are close but distinct of that of chrysotile(Fig. 4) as already described by Cathelineau et al. (2015) forthe whole solid solution.

The Ni-Mg kerolite-pimelite series from type 1 ore exhibitsstrong chemical zonation due to Ni-Mg substitutions. The Nicontents cover a large range from 10 wt% NiO to 27 wt% NiO.Type 2 ores, conversely, display two narrow ranges of Ni con-centrations recorded with rather constant concentrations(Fig. 4e, f): NiO ranges from 0 to 0.5 wt% in the white Mg-kerolite and from 43 to 49 wt% in the Ni-kerolite (pimelite).The two ranges are close to those of the end-member values forthe Ni-Mg kerolite solid solution. All the analysed kerolitesshow a significant stoichiometric deviation from talc, given thatan excess in the octahedral Ni (+Mg) occupancy is observed incorrelation with a deficiency in the tetrahedral occupancy.

From all observations, it can be concluded that these min-erals are rather distinct both from serpentine or talc and are notmixtures with any of these minerals. Although IMA does notrecognize kerolite as a mineral phase, the term kerolite is usedbecause this phase is distinct from talc, is not mixed withserpentine and has a rather narrow range around 0.50±0.03for the Si/(Si + Mg + Ni) ratio, distinct from both that of talc(0.57) and serpentine (0.4) (Fig. 5). The value of this ratio hasbeen largely discussed in the literature: Brindley and Hang(1972) explain this deviation either by the presence of anadditional brucite layer or by a break in the tetrahedral sheet.Another hypothesis is the physical mixing with serpentinelayers suggested for the Dominican ores by Villanova-de-Benavent et al. (2014). This mixing justifies, in that particularcase, the use of the term garnierite, e.g., a mechanical mixingbetween serpentine, talc-like and sepiolite minerals. In ourcase, the deviation cannot be easily attributed to the presenceof serpentine layers given that serpentine layers have been notobserved under TEM and that the Si/(Si + Mg + Ni) ratio isconstant with no chemical trend between kerolite and serpen-tine as that found on Dominican ores by Villanova-de-Benavent et al. (2014). Thus, the term garnierite cannot beused as the studied minerals are not a mix of talc-like andserpentine with a fixed abundance for the two phases.

Discussion

Field observations and the laboratory characterization indicatethat two types of Ni silicate ores occur in the Koniambomassif

Table 1 Average composition of kerolite from veins (ore type 1) andtarget-like features (ore type 2) of the analyses presented in Fig. 5,measured by EPMA

Type 1 Type 2 (target-like ore)

Bluish kerolite Pimelite (green) Kerolite (white)

wt% Av σ Av σ Av σ

SiO2 52.09 1.89 38.49 1.80 52.25 3.10

Al2O3 0.01 0.01 0.03 0.05 0.02 0.03

MgO 23.98 2.71 2.27 1.15 35.92 1.42

K2O 0.06 0.05 0.00 0.00 0.00 0.00

Cr2O3 0.00 0.01 0.00 0.00 0.00 0.00

MnO 0.02 0.03 0.02 0.04 0.04 0.13

FeO 0.02 0.02 0.12 0.14 0.41 1.16

CoO 0.07 0.06 0.44 0.31 0.04 0.06

NiO 14.31 4.28 49.65 1.44 0.10 0.11

Total 90.55 91.01 88.76

n 50 8 20

% NiO min 10.29 46.63 0.00

% NiO max 27.33 51.06 0.37

Si 3.78 3.50 3.62

Al 0.00 0.00 0.00

∑ Oct. 3.78 3.51 3.62

Fe2+ 0.00 0.01 0.02

Mg 2.59 0.31 3.72

Ni 0.84 3.64 0.01

∑ Oct. 3.43 3.95 3.75

Miner Deposita (2016) 51:271–282 277

Page 143: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

near the oxidation zone. They are easily discriminated by thefollowing features:

– Veins distributed as clusters of sub-parallel fractureinfillings with a predominant strike and dip, at leastat the open pit scale. The syntectonic characteristicis shown by the existence of hydraulic breccia andstriated planes affecting both the silicate infillingsand the fracture walls (Cluzel and Vigier 2008).They can be observed over a large depth range offew tens meters to few hundred meters below pres-ent day surface. Each vein is characterized by a

large range of Ni-Mg content with complex chemi-cal zoning. They are commonly associated with red-dish micro-crystalline quartz.

– Target-like ores consist of a thin layer cementing a jointgenerally less than 0.5 mm in thickness, where the jointdoes not have a preferred orientation. These joints areinterpreted to result from decompression that is typicalof the first hundreds of meters below the surface; theyare not syntectonic. The target-like ores are exclusivelyobserved in a narrow zone of a few tens of meters inthickness, just below the high-oxidation zone in thesaprolitic horizon.

Fig. 4 a, b Samples of target-like(ore type 2) analysed by RAMAN(locations are shown by whitedots). b, c Zoom on the Mg-kerolite and Ni-kerolite (pimelite)zones of a target-like. e, fBack-scattered electron image ofthe botryoidal habit of Ni-kerolite(f) and Mg-kerolite (e). Eachgrowth band has a rather constantNi concentration characterized byits specific mean Z (averageatomic weight). Mg-Tl, Ni-Tl andSp are respectively Mg talc-like,Ni talc-like and serpentine. TheNiO concentrations are in wt%and were determined by electronmicroprobe analyses. RAMANspectrum representative of theMgand Ni kerolite from target-like(bottom on the figure)

278 Miner Deposita (2016) 51:271–282

Page 144: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Time-space conditions favourable to target-like oreformation

The specific hydrologic conditions presently affecting thehighest parts of the peridotite relief induce strong water-rockinteractions characterized by a high rate of water infiltration,and subsequently, nearly continuous hydrolysis of all mineralphases except insoluble minerals (Fe-oxihydroxides) andinherited low solubility oxides (e.g., chromite) under oxidizingconditions. The almost complete dissolution of the serpentinizedperidotite along the faults results in large open dissolution pipes,which are sub-vertical to dipping 60–70°. These large pipes arelocally collapsed and cemented by white quartz and pimelite, asobserved in breccia from the Cagou pit (Fig. 2a, b). These pro-cesses are rather distinct from those classically described for theslow downward progress of the nearly horizontal laterite fronts(Nahon and Tardy 1992; Freyssinet and Farah 2000).

All the observed features of the type 2 ore indicate that itformed under supergene conditions, close to the present-daytopographic surface, and therefore during recent times. Target-like ores can develop once the very late exhumation of theperidotite massif and its erosion has induced mechanical de-compression and subsequent jointing. The bulk permeabilityof the first hundreds of meters below the surface is thus in-creased by this process (Hencher et al. 2011). The high perme-ability of the saprocks allows rapid vertical transfer and shortresidence time for the incoming waters. Spatial relationshipsbetween the topographic surface and the occurrence of target-like ores around several tens of meters (30 to 50 m) below thissurface suggest the potential link with the oscillating level of thewater table. In this zone, the water table undergoes significantoscillations which are linked to the seasonal variations of waterprecipitation. In addition, the topology of the boundary of thewater saturated zone is controlled by local hydrological condi-tions which primarily depend on the relief, here characterized

by a hilly landscape. Target-like ores are therefore quite distinctfrom the ores that formed during fault activity.

Physical-chemical processes linked to target-like oreformation

The formation of target-like ores implies a mechanism of min-eral precipitation from interstitial water present in the joints,e.g., an increase in (Mg2+)/(H+)2 and (Ni2+)/(H+)2 for a givenpH (Fig. 6). Water close to the topographic surface is far from

Fig. 5 Ternary Si-Mg + Fe-Ni diagram with analytical results for thekerolite veins (Ni-Mg solid solution, ore type 1) and target-like (Mgand Ni-kerolite end-members, ore type 2)

Fig. 6 Conceptual model of the Mg-Ni-kerolite target-like formation:Top cross-section peridotite block limited by joints, some partially filledwith early type 1 ores, acting as a fluid pathway and affected bydissolution. Infiltration into the rather closed fracture occurs mostlythrough diffusion; fracture and sequence of target-like precipitation with(1) partial dehydration during an episode of water table lowering, withprecipitation of Ni-kerolite (2) at the margins, and Mg-kerolite (3) at ahigher degree of fissure dehydration (bottom)

Miner Deposita (2016) 51:271–282 279

Page 145: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

mineral saturation, due to fast infiltration and a short residencetime, thereby preventing any saturation with respect to sili-cates. The oversaturation of the solution with respect to thesilicates therefore requires an additional mechanism thatwould increase cation activities in solution. The most likelymechanisms are (i) an increase in the residence time of thewater, away from the main drainage open fractures, whichwould increase the rate of interaction with rocks when inter-stitial water is trapped in the micropores of the closed joints, aprocess which results in an increase in the cation/H+ ratio dueto the hydrolysis of pre-existing minerals, in rocks and frac-tures, and (ii) water evaporation, which occurs several timesper day in between the rain events, in the rock volume locatedabove the saturated zone. In this volume, interstitial water isconnected with a major dissolution zone and drainage pipesand may be enriched in Ni, from the dissolution of all Mg (Ni)silicates including earlier ores, such as type 1 ores. Thus, theevaporation of waters containing Si, Mg and Ni may result inthe saturation of pimelite or Mg-kerolite. The mineral distri-bution at the origin of the spectacular target-like colour zoningseems to come from chromatographic process due to the sat-uration of pimelite at the edge and saturation, with respect toMg-rich kerolite, in the subsequent stages of the evaporationprocess at the centre of the block joint. It is suggested that thelast zone to dry is in the centre of the joint, and there, waterprecipitates the remaining Si and Mg once Ni is already pre-cipitated as pimelite near the joint boundary, or alternately,Mg-rich silicate could also have precipitated earlier. Pimeliteis much more insoluble than Mg-kerolite, thereby explainingwhy this mineral precipitates first despite a Ni/Mg ratio inwater close to 10−3; this process was already described byGalí et al. (2012). The process is probably recurrent and re-peated several times before reaching a significant volume of

precipitated minerals but is limited to joints with a sufficientlylow permeability, preventing rapid dissolution and the releaseof elements into the percolating waters. It is interesting to notethat under these conditions, only the end-members of thekerolite-pimelite solid solution can precipitate. In open joints,the rate of infiltration is such that it prevents precipitation orfavours mostly dissolution of all silicates.

The effects of water infiltration on the pre-existing target-like mineral are visible on the blocks exposed in the quarries.After a period of several months, the nice green glass translu-cent pimelite micro-spherolites are replaced by pulverulentamorphous phases resulting from pimelite hydrolysis, whichare difficult to identify. Consequently, a large part of the nickelbearing joints are only revealed during mechanicalfracturation during open pit quarrying, and this explainswhy, in the other joints or faults, most of the Ni-bearing min-erals are partly dissolved.

Transfers at the bottom of the saprolitic horizon

The high number of open joints in the first several hundredmeters of the peridotite, in addition to the high drainage zonessuch as the dissolution pipes, allows an efficient vertical trans-fer of water towards the water saturated zone (Fig. 7). Zonescharacterized by rapid vertical water transfer lack mineral pre-cipitation. They are mostly characterized by residual ironoxyhydroxides, the trivalent iron being immediatelyreprecipitated in situ. Conversely, less open joint networksmay be affected by the reprecipitation of Ni-Mg silicates suchas the kerolite-pimelite series, mostly when located near clus-ters of early type 1 ore fractures which act as a source of Ni.Thus, pimelite occurs in a variety of joints; their distribution israther distinct from the syntectonic kerolite fractures,

Fig. 7 Redistribution of the earlyores located in discontinuities(ore type 1 as kerolite) at theinterface between the oxidizedand reduced zones within thewater table oscillation zone.Newly formed ores (ore type 2 aspimelite) occur as target-likepatterns in joint thin layers. Thisredistribution is independent ofthe occurrence of laterite, whichformed much earlier

280 Miner Deposita (2016) 51:271–282

Page 146: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

confirming that it represents a later Ni-precipitation in thejoints that is genetically distinct from the earlier type 1. Theabundance of type 2 ore pimelite joints at the bottom of theregolith suggests that type 2 ores significantly contribute to thegrade of the main ore type exploited today in the open pits,besides Ni-rich iron oxyhydroxides, whereas type 1 ores indiscontinuities only represent isolated and rich clusters, e.g.,a very small portion of the exploited ores. As a result, type 1ores are not considered when estimating the reserves due totheir scarcity and erratic distribution, as well as to their veryhigh, but not representative, Ni-grades.

Conclusion

Two types of silicate ores are distinguished in the saprolitefrom the Koniambo Massif:

– An early syntectonic mineralization, in relation with ma-jor discontinuities or sets of fractures, producing high-grade ores dominated by Mg-Ni kerolite-pimelite solid-solution phases.

– A later Ni silicate mineralization in narrow joints occur-ring as rims of pimelite around target-like joint fillingsand produced by present-day or recent shallow supergeneprocesses. The latter are favoured by the infiltration ofmeteoric waters in the decompressed bedrock and over-laying saprolite and subsequent water-rock interactions.The crystallization of pimelite at the bottom of the sapro-lite may explain the presence of the silicates found at thebase of the oxidized saprocks, generally difficult to iden-tify when mixed with iron oxyhydroxides.

The target-like ores result from a redistribution process.This process seems to be very recent, if not still active, andlinked to the present-day topography and water table move-ments. It is probably independent of the main stages of lateriteformation, which are thought to have occurred in earlier times,during the Cenozoic. This pervasive redistribution of nickel islikely responsible, in large part, for the rather large Ni anom-alies within the saprocks, which make up the bulk of theexploited ore and for the rather monotonous Ni-grade withinthe saprolite, ranging from 2 to 3 wt%. Conversely, the type 1ores, exclusively located in discontinuity networks, are re-sponsible for high but erratic grades. However, the Ni-enriched zones are partly controlled by the early distributionof the type 1 fractures. This tends to indicate that the geometricdistribution of type 1 Ni ores may have influenced the present-day distribution of exploitable ores.

From the mining point of view, this significant redistribu-tion of the early concentrations is of primary importance fordetermining the mining volumes that can be exploited. Thus,the mineralized domains within the first hundred meters are

far more extensive than the syntectonic type 1 kerolite fracturenetworks which constitute the richest ores exploited in NewCaledonia during the very first stages of exploitation at the endof the nineteenth and the beginning of the twentieth century. Itcan be concluded that the main consequence of the redistribu-tion of the earlier type 1 Ni-concentration is particularly inter-esting for mining as the large lateral redistribution ensures arather homogeneous volume of mineralized rock.

Acknowledgments This work has been carried out thanks to the finan-cial and technical help from Koniambo S.A. and Labex Ressources 21(supported by the French National Research Agency through the nationalprogram.

BInvestissements d’avenir^, reference ANR-10-LABX-21–LABEXRESSOURCES 21), with analytical contributions from the UMRGeoRessources No 7356 laboratory platforms. The field work was basedon an early introduction to this subject matter during a first 1-day visit toKoniambo in 2011 by C. Couteau and E. Fritsch (field trip organized aspart of a CNRT project (French National Centre for Technological Re-search: BNickel and its environment^) and mostly thanks to 2-week staysboth in 2012 and 2013 financed by Koniambowithin the framework of B.Quesnel’s PhD thesis. Sara Mullin, a translator specializing in scientificpublications, improved the English. We thank the associate editor Thom-as Bissig, the editor-in-chief George Beaudoin, and an anonymous re-viewer for their constructive reviews which allowed to improve thispaper.

References

Brindley GW, Hang PT (1972) The hydrous magnesium-nickel silicateminerals (so-called garnierites). Int Clay Conf 1:41–50

Brindley GW, Hang PT (1973) The nature of garnierites—I structures,chemical compositions and color characteristics. Clay Miner 21:27–40

Brindley GW, Maksimovic Z (1974) The nature and nomenclature ofhydrous nickel-containing silicates. Clay Miner 10:271–277

Brindley GW, Wan HM (1975) Compositions, structures, and thermalbehaviour of nickel-containing minerals in the lizardite–nepouiteseries. Am Mineral 60:863–871

Brindley GW, Bish D, Wan HM (1977) The nature of kerolite, its relationto talc and stevensite. Mineral Mag 41:443–452

Brindley GW, Bish D, Wan HM (1979) Compositions, structures, andproperties of nickel-containing minerals in the kerolite–pimelite se-ries. Am Mineral 64:615–625

Butt CRM, Cluzel D (2013) Nickel laterite ore deposits: weatheredserpentinites. Elements 9:123–128. doi:10.2113/gselements.9.2.123

Cathelineau M, Caumon MC, Massei F, Brie D, Harlaux M (2015)Raman spectra of Ni-Mg kerolite: effect of Ni-Mg substitution onO-H stretching vibrations. J Raman Spectrosc, ProceedingsGEORaman 2014 (11th), ST Louis (USA). doi:10.1002/jrs.4746

Chevillotte V, Chardon D, Beauvais A, Maurizot P, Colin F (2006) Long-term tropical morphogenesis of New Caledonia (Southwest Pacific):importance of positive epeirogeny and climate change.Geomorphology 81:361–375

Cluzel D, Vigier B (2008) Syntectonic mobility of supergene nickel oresof New Caledonia (Southwest Pacific). evidence from garnieriteveins and faulted regolith. Resour Geol 58:161–170. doi:10.1111/j.1751-3928.2008.00053.x

Cluzel D, Aitchison JC, Picard C (2001) Tectonic accretion and under-plating of mafic terranes in the Late Eocene intraoceanic fore-arc of

Miner Deposita (2016) 51:271–282 281

Page 147: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

New Caledonia (Southwest Pacific). geodynamic implications.Tectonophysics 340(1–2):23–59

Cluzel D, Jourdan F, Meffre S, Maurizot P, Lesimple S (2012) The meta-morphic sole of New Caledonia ophiolite: 40Ar/39Ar, U-Pb, andgeochemical evidence for subduction inception at spreading ridge.Tectonics 31:TC3016. doi:10.1029/2011TC003085

Dublet G (2012) Relation entre spéciation et distribution du nickel dansles couvertures d’altération latéritique des roches ultrabasiques deNouvelle-Calédonie. PhD thesis, UPMC Paris VI, 353 p

Dublet G, Juillot F, Morin G, Fritsch E, Fandeur D, Ona-Nguema G,Brown GE Jr (2012) Ni speciation in a New Caledonian lateriticregolith: a quantitative X-ray absorption spectroscopy investigation.Geochim Cosmochim Acta 95:119–133

Fandeur D (2009) Géochimie et cristallochimie du chrome au cours del’altération de roches ultrabasiques en Nouvelle-Calédonie (Massifdu Koniambo). PhD Thesis, Université Paris Diderot–IPGP, 321 p

Freyssinet P, Farah AS (2000) Geochemical mass balance and weatheringrates of ultramafic schists in Amazonia. Chem Geol 170:133–151

Freyssinet P, Butt CRM, Morris RC, Piantone P (2005) Ore formingprocesses related to laterite weathering. Economic Geology, 100thAnniversary Volume: 681–722

Galí, Soler JM, Proenza JA, Lewis JF, Camas, Tauler (2012) Ni-enrichment and stability of Al-free garnierite solid- solutions: a ther-modynamic approach. Clays Clay Miner 60:121–135

Genna A, Maurizot P, Lafoy Y, Augé T (2005) Contrôle karstique deminéralisations nickélifères de Nouvelle-Calédonie. ComptRendus Geosci 337:367–374

Gerard P, Herbillon AJ (1983) Infrared studies of Ni-bearing clay min-erals of the kerolite–pimelite series. Clays Clay Miner 31:143–151

Gleeson SA, Butt CRM, Elias M (2003) Nickel laterites: a review. SocEcon Geol Newslett 54:9–16

Hencher SR, Lee SG, Carter TG, Richards LR (2011) Sheeting joints:characterization, shear strength and engineering. Rock Mech RockEng 44:1–22

Manceau A, Calas G (1985) Heterogeneous distribution of nickel in hy-drous silicates from New Caledonian ore deposits. Am Mineral 70:549–558

Manceau A, Calas G (1986) Nickel-bearing clay minerals: IIIntracrystalline distribution of nickel: an X-ray absorption study.Clay Miner 21:341–360

Manceau A, Calas G, Decarreau A (1985) Nickel-bearing clay minerals:I. Optical spectroscopic study of nickel crystal chemistry. ClayMiner 20:367–387

Manceau A, Schlegel ML, Musso M, Sole VA, Gauthier C, Petit PE,Trolard F (2000) Crystal chemistry of trace elements in natural andsynthetic goethite. Geochim Cosmochim Acta 64(21):3643–3661

Maurizot P, Lafoy Y, Poupée M (2002) Cartographie des formationssuperficielles et des aléas mouvements de terrain en Nouvelle-Calédonie, Zone du Koniambo: Bureau de RecherchesGéologiques et Minières, Public Report RP51624-FR, 45 p

Maurizot P, Vendé-Leclerc M, Collot J (2009) Carte géologique de laNouvelle-Calédonie au 1/500.000. Service Géologique deNouvelle-Calédonie/DIMENC-BRGM, Nouméa

Nahon DB, Tardy Y (1992) The ferruginous laterites. In: Butt CRM,Zeegers H (eds) Regolith exploration geochemistry in tropical andsubtropical terrains, vol 4, Handbook of Exploration Geochemistry.Elsevier, Amsterdam, pp 41–55

Paquette J-L, Cluzel D (2007) U–Pb zircon dating of post-obductionvolcanic-arc granitoids and a granulite-facies xenolith from NewCaledonia. Inference on Southwest Pacific geodynamic models.Int J Earth Sci 96:613–622. doi:10.1007/s00531-006-0127-1

Paris JP (1981) Carte géologique de la Nouvelle Calédonie, feuille Ouest,1/200000. Bureau de Recherches Géologiques et Minières (BRGM)

Pelletier B (1983) Localisation du nickel dans les mineraisBgarnieritiques^ deNouvelle-Calédonie. Sci GeolMem73:173–183

Pelletier B (1996) Serpentines in nickel silicate ore from New Caledonia.In: Grimsey EJ, Neuss I (eds) Proceedings, nickel 096: mineral tomarket, Kalgoorlie. The Australasian Institute of Mining andMetallurgy, Carlton, pp 197–205

Printemps J, Ausseil A-G, Dumas P, Mangeas M, Dymond JR, Lille D(2007) An Erosion Model for Monitoring the Impact of Mining inNew Caledonia. MODSIM conference

Schellman W (1983) Geochemical principles of lateritic nickel ore for-mation. In: Melfi AJ, Carvalho A (eds) Proceedings of the IIInternational Seminar on Lateritisation Processes. Sao Paulo,Brazil, 1982, 119–135

Sevin B, Ricordel-PrognonC, Quesnel F, Cluzel D, Lesimple S,MaurizotP (2012) First palaeomagnetic dating of ferricrete in NewCaledonia:new insight on themorphogenesis and palaeoweathering of BGrandeTerre^. Terra Nov. 24(1):77–85

Springer G (1974) Compositional and structural variations in garnierites.Can Mineral 12:381–388

Terry JP, Wotling G (2011) Identifying rain-shadow influences on riverflows and flood magnitudes across the central massif divide of LaGrande Terre Island, New Caledonia. J Hydrol 404:77–86

Trescases JJ (1975) L’évolution géochimique supergène des rochesultrabasiques en zone tropicale. Mem ORSTOM 78: 279 p

Troly G, Esterle M, Pelletier B, Reibell W (1979) Nickel deposits in NewCaledonia, some factors influencing their formation. In: Evans DJIet al (eds) International Laterite Symposium. New Orleans,Louisiana, 19–21 February. Society of Mining Engineers, NewYork, pp 85–119

Villanova-de-Benavent C, Aiglsperger T, Jawhari T, Proenza JA, Galí S(2012) Micro-Raman spectroscopy of Garnierite minerals: a usefulmethod for phase identification. Rev Soc Esp Miner 180:181–16

Villanova-de-Benavent C, Proenza JA, Galí S, García-Casco A, Tauler E,Lewis JF, Longo F (2014) Garnierites and garnierites: textures, min-eralogy and geochemistry of garnierites in the Falcondo Ni-lateritedeposit, Dominican Republic. Ore Geol Rev 58:91–109

Wells MA, Ramanaidou ER, Verrall M, Tessarolo C (2009) Mineralogyand chemical chemistry of Bgarnierites^ in the Goro lateritic nickeldeposit, New Caledonia. Eur J Mineral 21:467–483

Xstrata (2012) Mineral resources and reserves at 31 Dec. 2012. AnnualReport, 50 p

282 Miner Deposita (2016) 51:271–282

Page 148: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 149: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genesis and mineralogy of Ni ores

Quesnel, B, Le Carlier de Veslud, C., Boulvais, P, Gautier, P, Cathelineau, M, Drouillet, M (2017)

3D modeling of the laterites on top of the Koniambo Massif, New Caledonia: refinement of the per descensum lateritic model

for nickel mineralization.

Miner Deposita.DOI 10.1007/s00126-017-0712-1

- 11 -

Page 150: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

ARTICLE

3D modeling of the laterites on top of the Koniambo Massif,New Caledonia: refinement of the per descensum lateritic modelfor nickel mineralization

Benoît Quesnel1 & Christian Le Carlier de Veslud1& Philippe Boulvais1 & Pierre Gautier1 &

Michel Cathelineau2& Maxime Drouillet3

Received: 17 February 2016 /Accepted: 5 January 2017# Springer-Verlag Berlin Heidelberg 2017

Abstract Resulting from the weathering of the PeridotiteNappe, laterites are abundant in New Caledonia and hostone of the largest nickel deposits worldwide. This work pre-sents a 3D model of the Koniambo nickel laterite ore deposit.It shows that the laterites are located along the ridges of themassif and organized as hectometric-sized patches obliquelycut by the topography and distributed at various elevations.

Three kinds of geometry were observed: (i) a thick lateritecover (between 20 and 40 m) overlying saprolite and mainlylocalized on topographic highs, (ii) a thin laterite cover (from afew meters to 20 m) mainly localized on areas with gentleslopes, and (iii) exposure of saprolite without laterite cover.Our data show that Ni-rich and Ni-poor areas are organized ashectometric-sized patches which broadly correlate with thedistribution of the laterite thickness. The highest Ni areas arelocalized on slopes where laterite cover is thin or absent. Theareas with lowest Ni are located in topographic highs underthe thickest laterite cover. The vertical Ni mass balance foreach borehole shows that, in areas with thick laterite cover,Ni is sub-equilibrated to slightly depleted whereas in areaswith thin laterite cover, Ni is enriched. This suggests the ex-istence of lateral infiltration of water rich in dissolved Ni, fromareas such as topographic highs to downstream slope areas, ina process leading to enrichment of saprolite in Ni in slopeareas. Mechanical transport and leaching of laterite materialon slopes, including Ni-bearingmaterial, could also contributeto local enrichment of Ni in the saprolite.

Introduction

Lateritic nickel deposits are associated with the intenseweathering of ultramafic rocks. Laterite development involvesdissolution of the primary minerals of the peridotite, whichleads to (i) leaching of soluble elements (Si, Mg) and (ii) insitu neoformation of mineral phases (mainly oxy-hydroxides)that host the insoluble elements (Fe, Al, and Cr). Nickel, withan intermediate behavior, is concentrated at the base of thelateritic profile, in the saprolite horizon, where it reaches eco-nomic concentrations (>1 wt.%). This is particularly the casein New Caledonia (Trescases 1975; Freyssinet et al. 2005) forwhich a typical lateritic profile is shown in Fig. 1. At the top,

Editorial handling: B. Orberger

Electronic supplementary material The online version of this article(doi:10.1007/s00126-017-0712-1) contains supplementary material,which is available to authorized users.

* Benoît [email protected]

Christian Le Carlier de [email protected]

Philippe [email protected]

Pierre [email protected]

Michel [email protected]

Maxime [email protected]

1 Géosciences Rennes - OSUR, Université Rennes 1, UMR 6118CNRS, 35042 Rennes Cedex, France

2 CNRS, CREGU, GeoRessources lab, Université de Lorraine, 54506Vandoeuvre-lès-, Nancy, France

3 Service géologique, Koniambo Nickel SAS, 98883Voh Nouméa, New Caledonia, France

Miner DepositaDOI 10.1007/s00126-017-0712-1

Page 151: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the laterite zone is mainly composed of iron oxy-hydroxides(hematite, goethite). Beneath the laterite horizon, the saprolitezone represents an intermediate state of alteration betweenlaterite and the fresh peridotite and hosts the most concentrat-ed Ni ores.

For the New Caledonian Ni laterite deposits, as forothers worldwide, a downward model of fluid circulationis classically invoked to explain the geometry of the lat-eritic profile and the distribution of elements (Glasser1904; de Chetelat 1947; Avias 1969; Trescases 1975;Fig. 1). According to this per descensum model, the su-pergene alteration of peridotite leads to vertical leachingand passive transport of Ni (and other mobile elements asMg and Si) through discontinuities crosscutting the peri-dotite. Nickel is then immobilized in the saprolite zone.Even if this model is a simple way to explain the Nilaterite deposit formation, several studies suggest renewedexamination of the per descensum model is warranted.First, evidence of low-temperature, hydrothermal condi-tions of formation for the silica veins commonly associat-ed with garnierite (Quesnel et al. 2016a; Fritsch et al.2016), of hydraulic brecciation involving silica and garni-erite (Cathelineau et al. 2016b), and of syn-tectonic for-mation of garnierite veins (Cluzel and Vigier 2008;Villanova-de-Benavent et al. 2014) suggests that the con-ditions of fluid infiltration leading to the irregular oreformation are more complex than simple downward per-colation of meteoric water. Second, evidence of current torecent Ni dispersion in the saprolite zone related to water

table movement (Cathelineau et al. 2016a) and geologicalobservations showing that Ni-rich areas are usually notspatially associated to the main laterite body (deChetelat 1947; Avias 1969; Trescases 1975, Golightly1981) suggest that formation of the Ni laterite ore deposithas to be thought of as a 3D process, implying lateral Nimobility.

In characterizing Ni mobility during laterization,most previous works discussed Ni behavior based onone or few isolated vertical profiles (e.g., Trescases1975; Fu et al. 2014) and rarely based on a significantsample dataset (Schellmann 1989). Restricted by thedifficulty in obtaining a large and regularly distributedborehole database, such studies can only be performedin 1D.

In 2007, Koniambo Nickel SAS initiated a large min-ing and industrial site for Ni production in theKoniambo Massif (Fig. 2). As a result, new outcropsand a large number of boreholes provide an exceptionaldataset to study the Ni laterite deposits. When investi-gating a complex geological case study, 3D geometricalmodeling coupled with field survey represents a first-order tool for integrating various data, building consis-tent geological structures, and performing quantitativeestimates. In this study, we combine 3D modeling withfield observations and a borehole dataset in order tocharacterize the distribution and the geometry of thelateritic surfaces and, more specifically, to estimatehow much lateral Ni mobility has contributed to theformation of Ni-rich laterite ore deposit.

Geological setting

Generalities

To date, nickeliferous laterites from New Caledonia host11% of the global reserves in nickel (USGS, U.S GeologicalSurvey 2016). Laterites are actively worked by several miningc om p a n i e s ( KN S , SMS P NMC , S LN , Va l eNouvelle-Calédonie, Montagnat, Ballande) making NewCaledonia the 6th nickel producer in the world (Baille et al.2013). The mainland of New Caledonia comprises a northernextension of the Norfolk Ridge and is approximately 50 kmwide and 400 km long. Ultramafic rocks are abundant,representing about 40% of the island surface. They mainlyconsist of peridotite and occur as a main massif calledBMassif du Sud^ in the southern part of the island and as aseries of klippes along the northwestern coast (Fig. 2). Theseperidotites overlie a volcano-sedimentary basement with asub-horizontal contact (Avias 1967; Guillon 1975). This ge-ometry resulted from the Eocene obduction of the NewCaledonia Peridotite Nappe on the Norfolk Ridge (Cluzel

Fig. 1 Typical laterite profile of New Caledonia developed on aperidotitic protolith (modified from Pelletier 1989 and Ulrich 2010)

Miner Deposita

Page 152: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

et al. 2001, 2012). The nappe is mostly composed ofharzburgite except in the northernmost klippes wherelherzolites dominate (Ulrich et al. 2010). In areas of highelevation in the Massif du Sud, the peridotites are locallyoverlain by layered pyroxenite and gabbro, interpreted to rep-resent the base of an oceanic crust (Guillon 1975; Prinzhoferet al. 1980; Paris 1981). In peridotite, compositional layeringis represented by 1- to 100-m-thick layers of dunite within themain harzburgitic mass (Ulrich et al. 2010). In the main part ofthe nappe, the degree of serpentinization of peridotite is vari-able but moderate (Orloff 1968; Quesnel et al. 2016b).Serpentinization becomes pervasive along the base of thenappe, forming a Bsole^ in which deformation has been in-tense (Avias 1967; Orloff 1968; Guillon 1975; Leguéré 1976;Cluzel et al. 2012; Quesnel et al. 2013).

Laterites are well developed at the top of each perido-tite massif. Some occurrences of pseudo-karstic structuresin the peridotite nappe demonstrate the existence of anefficient drainage system (Trescases 1975; Genna et al.2005). Laterites occur as several planation surfaces at dis-tinct elevations. They probably did not form in a singleevent. On the northwestern klippes, the planation surfacesare interpreted as resulting from post-obduction

progressive uplift of the island coupled with events offavorable climatic conditions (Latham 1977, 1986;Chevillotte et al. 2006). For the Massif du Sud, someauthors consider that uplift and climate also explain theformation of several planation surfaces (Chevillotte et al.2006), whereas others consider that this geometry wasformed, at least in part, by late brittle tectonic events(Trescases 1975; Latham 1977). Most of these surfacespresent evidence of reworked material (Trescases 1975;Latham 1977, 1986) attesting to alternating periods ofwet and warm climate, which led to laterization, and drierperiods, which led to erosion (Latham 1977; Chevillotteet al. 2006). Dating these surfaces is still a challenge;however, the older surfaces probably formed before ca.20 Ma (Latham 1986; Chevillotte et al. 2006; Sevinet al. 2012; Quesnel et al. 2013).

The KoniamboMassif is one of the klippes of the PeridotiteNappe located along the west coast of the main island (Fig. 2).It is 20 km long by 5 km wide (Fig. 3), and it comprisesharzburgite with interlayers of dunite that define a crude com-positional layering (Maurizot et al. 2002; Ulrich 2010). TheKoniambo Massif rises from a narrow coastal plain to 940 mabove sea level. The degree of serpentinization increases from

Fig. 2 Simplified geologicalmap ofNewCaledonia.Ultramafic rocks are fromMaurizot andVendée-Leclerc (2009), and laterites are adapted fromParis (1981)

Miner Deposita

Page 153: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the top to the base of the massif (Maurizot et al. 2002) with (i)a massive serpentinite level forming the base of the massif, (ii)a highly serpentinized level forming the intermediate part ofthe massif, and (iii) a moderately serpentinized and highlyfractured level (Leguéré 1976) forming the upper part of themassif on which laterite is developed. This massif is known tocontain significant, high-grade nickel laterite mineralizations(Xstrata 2012). The laterites are organized as distinct planationsurfaces, highly dissected and partly reworked, localized atelevations above ∼400 m (Fig. 3a, Maurizot et al. 2002). Atlower elevations, laterites of the westerly inclined KaféatéPlateau probably belong to a younger planation surface(Latham 1977; Chevillotte et al. 2006).

Alteration facies

Figure 1 shows the typical lateritic succession of NewCaledonia, characteristic of laterites developed under wet cli-mate and tectonically active terranes with moderate relief(Golightly 1981). The upper part of the profile is composedof (i) the iron duricrust and (ii) the laterite horizon, composedof iron oxy-hydroxides (mainly goethite and minor hematite)and separated in two zones as a function of the alteration. Thered laterite zone results of a higher level of weathering of theperidotite than the underlying yellow laterite zone. Under thelaterite horizon, the saprolite consists of a zone with interme-diate state of alteration between parent rock and laterite. The

Fig. 3 a Geological map of the Koniambo Massif adapted from Maurizot et al. (2002) draped on the 10-m digital elevation model. C1 and C2 are thelocalization of the two cross sections in Fig. 5. b Localization of the borehole dataset

Miner Deposita

Page 154: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

earthy saprolite hosts the major part of the diffuse Ni miner-alization related to oxy-hydroxides where goethite is the mainNi carrier (Dublet et al. 2012). Immediately below, the rockysaprolite, in which the degree of alteration is lower and thestructure of the parent rock is partially preserved, hosts either apervasive Ni ore, where both silicates and oxy-hydroxides arethe Ni carriers (Dublet et al. 2012), and an erratic fracture-controlled high-grade Ni ore, known as garnierite. The termgarnierite classically refers to a mixture of several Ni-Mgphyllosilicates, although the main Ni silicate in veins fromthe Koniambo Massif is a Btalc-like phase^ (Cathelineauet al. 2016a). A recent study of fracture samples filled withgarnierite and of films in joints occurring as target-like oresfrom the Koniambo Massif revealed that nickel was essentiallycontained in the talc-like phases of the kerolite-pimelite solidsolution (Cathelineau et al. 2016a). This is coherent with pre-vious work on the Dominican Republic (Villanova-de-Benavent et al. 2014) and Venezuela (Soler et al. 2008) nickellaterite deposits.

It is a common impression, gained from the literature, thatlaterite facies occur as a regular succession of layers. However,two main types of configurations can be recognized in the field.The first, localized on topographic highs, is a thick laterite cover,often preserved as complete profiles, whereas the second is pre-served as thin and incomplete profiles localized on gentle slopes(de Chetelat 1947 his Fig. 7; Trescases 1975; Avias 1978). Inaddition, boreholes indicate that the succession of laterite facies ishighly irregular, both horizontally and vertically. Therefore, onlythe two main horizons were considered in this study: the upperlaterite zone and the lower saprolitic zone (Fig. 1).

Data and methodology

Data

The input data were mainly provided by the Koniambo NickelSAS Company. They include the following:

& A digital elevation model (DEM) with a horizontal reso-lution of 10 m

& A database of 6700 boreholes drilled in laterite, mainlyalong the ridges of the massif. The well head altituderanges from 325 to 925 m, but 50% of them lie in therange 675–815 m (Fig. 3b). For each borehole, analysesof major and some trace elements, lateritic facies, somespecific features (e.g., presence of garnierite), and densityare available on samples taken approximately every meter,altogether accounting for 260,000 data points. The maxi-mum depth reached by these boreholes is 120 m, but inaverage, the depth is about 30 m.

& The geological maps published in literature or from theBRGM

Two important points have to be noted. First, the boreholesare not regularly distributed, with spacing between adjacentboreholes ranging from a few hundred meters to 5 m. As theboreholes are concentrated along the ridges, the lower parts ofthe laterites, at intermediate altitude, are poorly represented ifnot sampled at all. Subsequently, some bias can occur duringinterpolation of data especially in areas where borehole den-sity is low or at the edges of the studied area due to boundaryeffects. On the contrary, the upper parts, corresponding to themining areas, are oversampled. Second, each chemical analy-sis corresponds to the mean value, measured on about 1 m ofthe sample composite, considered by the logging team as rep-resentative homogeneous material in a borehole. The chemicalcompositions have been obtained using X-ray fluorescencespectrometry. Thus, as no mineralogical or petrological studyis available on these samples, this kind of analysis is lessaccurate than any detailed study based on a limited numberof samples (Dublet et al. 2012, 2015), but it is well suited forlarge-scale mass balance. In this study, we focus mainly onanalyses of Ni, Al2O3, Fe2O3, Cr, MgO, and SiO2 in percent-age by weight. Loss on ignition (LOI) measured by calcina-tion ranged from 0 to 20%, with an average of 12%.

D modeling methodology

We used the GOCAD 3D numerical modeler (Mallet2002) to reconstruct the geometry of the lateritic coverand to quantify the distribution of geochemical tracers,including mass balance, in 3D. The GOCAD modeler pro-vides a discrete representation of geological objects interms of regular meshes (grids) or irregular meshes(polygonal curves, triangulated surfaces; Mallet 2002).Triangulated surfaces appear to be particularly relevantfor representing complex surfaces with irregularly distrib-uted data points, which is the case for the present study.The modeling process included four steps. First, a 3Ddatabase was built integrating all the available data. Ageo-referenced digital terrain model (DTM) consisting ofgeological maps draped over the DEM was constructed(Fig. 3a). Second, surface maps and the lithological andgeochemical information from the boreholes were integrat-ed. Third, the 3D reconstruction of geological structuresfrom borehole data, namely, the laterite/saprolite interface,was performed by discrete smooth interpolation (DSI;Mallet 2002) in order to study the geometry of the lateritecover. Finally, various computations were performed,based on the previous 3D objects, in order to characterizethe Ni distribution in the Koniambo Massif, including thefollowing:

& Study of the spatial distribution of elements with a specialfocus on the lateral variability of the Ni content in thesaprolite zone

Miner Deposita

Page 155: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

& Construction of isocon diagrams (Grant 1986) based onseveral representative boreholes in order to study the ver-tical behavior of elements during weathering

& Calculation of 1D vertical mass balance of Ni on thewhole borehole dataset in order to study the lateral Nidistribution at the scale of the Ni laterite ore deposit

Results

Geometry of the laterite cover

The laterites are present along the ridges of the massif and onthe gentle slopes located (typically <20°) close to the ridges.In contrast, almost no laterite is found along the steep slopes orin the valleys (Fig. 3a). The largest and most continuous lat-eritic horizon occurs over the northwest-trending axial ridgeof the massif. It is also in this area where most of the Nideposits are mined (Fig. 3b). At the scale of the massif, thelateritic horizons define a broad surface with a slope of about3° toward the SW. Some patches depart from this trend, oc-curring beneath this surface.

A 3D surface corresponding to the laterite-saproliteboundary was modeled using borehole data. Two crosssections (Fig. 4(c1, c2)), one parallel and one perpendicu-lar to the main NW-SE ridge axis, show the relationshipsbetween the laterite-saprolite interface and the topography.This interface is particularly interesting since it is knownthat most of the Ni mineralization is localized in the sap-rolite zone. It should be noted that the level of detail of thissurface is conditioned by the spatial density of boreholes.Globally (Fig. 4a, b), the laterite cover is organized ashectometric-sized patches obliquely cut relief and distrib-uted at various elevations (Fig. 4(c1, c2)). At first order,the interface between laterite and saprolite is parallel totopography (Fig. 4(c1, c2)) and may be extrapolated fromone patch to the neighboring one suggesting a past conti-nuity between them. Three kinds of configurations are ob-served. The first consists of topographic highs where thelaterite cover is thickest, ranging between 20 and 40 m.The second consists of gently sloping areas where lateritecover is thin (from a few meters to 20 m), commonly lo-cated downstream of the topographic high with thick later-ite cover. The third consists of areas without laterite cover,allowing the saprolite to be exposed at the surface as aresult of localized erosion on steep slopes.

In places where the density of boreholes is high, the surfacehas a complex multiscale shape with two kinds of features: (i)multiple small-scale depressions of metric to decametric verti-cal amplitude (Fig. 5a, b) and (ii) larger horseshoe-shaped de-pressions (Fig. 5c).

Spatial distribution variation of chemical composition

Composition of laterite facies

Most of the boreholes intersected unweathered serpentinizedperidotite at variable depths from a few meters to 120 m. Inorder to rule out a possible variability of bedrock, and partic-ularly for the Ni content which could lead to local variations inthe laterite Ni content, the geochemical homogeneity of thebedrock was investigated. The initial Ni content of unweath-ered peridotite is 0.2–0.4 wt.% in New Caledonia (Trescases1975; Audet 2008; Dublet et al. 2012). Consequently, a Nicontent greater than 0.5 wt.% is considered anomalous forNewCaledonia peridotite. Therefore, the bottom of each bore-hole was examined, discarding those that displayed a Ni con-tent >0.5 wt.%. Statistical analysis (1st and 3rd quartile,Table 1) indicates that the bedrock was homogeneous for themajor elements (Table 1). It should be noted that the peridotiterocks sampled here are a few hundred meters above the main,sub-horizontal contact of the peridotite nappe with the under-lying Poya unit and, thus far, from serpentinization associatedwith this contact. In comparison to the bedrock (Table 1), thesaprolite presents similar SiO2, higher Fe2O3, Al2O3, and Niand lower MgO contents. In the laterite, both SiO2 and MgOcontents are significantly lower whereas Fe2O and Al2O3 arehigher. Ni content is also higher, but less than in the saprolite.

Vertical variation of chemical composition

Three boreholes (Fig. 6 and Online Resource 1) were selectedas representative of the three profiles previously defined: well1—thick laterite, well 2—no laterite cover, and well 3—thinlaterite. In the laterite cover (wells 1 and 3), the concentrationof elements of interest is constant. Relative to the peridotiteprotolith, the oxide laterite displays high contents of Fe2O3

(70% wt), Al2O3 (5 wt.%), and Cr (4 wt.%) and low contentsof MgO (∼1 wt.%) and SiO2 (∼3 wt.%). The Ni content in-creases toward the interface with saprolite, where its concen-tration locally reaches 2 wt.%. In the saprolite zone, elementconcentrations are variable. In wells 1 and 2, SiO2 is near 35–40 wt.%, MgO is 25–40 wt.%, Fe2O3 is 5–15 wt.%, Al2O3 is1 wt.%, and Cr is 0.7 wt.%. Ni is also variable but enrichedrelative to peridotite with a concentration of 1–3 wt.%. In well3, under a thin laterite cover, the saprolite is slightly higher inSiO2 (45 wt.%) and significantly higher in Ni (3 wt.%). TheMgO content decreases in the first 30 m and then increaseswith a mean value around 20 wt.% for the saprolite interval.Fe2O3 shows the opposite trend with a small increase in thefirst 30 m with a mean concentration of 15 wt.% Fe2O3 andthen decreases to a mean value of 10 wt.%. In well 3, MgOand Ni are positively correlated whereas they are anti-correlated in the other wells (1 and 2). This likely reflectsthe presence of garnierite, consistent with the erratic, higher

Miner Deposita

Page 156: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

values of Ni. In wells 1 and 2, the positive correlation betweenNi and Fe2O3 indicates that goethite is the main Ni carrier, asdemonstrated by Dublet et al. (2012) in the same massif.

A comparison of the geochemical features of samples fromthe three wells is presented in a ternary plot (Fig. 7). Severalgeochemical trends are observed for well 1. The main trendcorresponds to the hydrolysis of Mg-(Ni) silicates yielding tocomplete loss of Si and Mg during laterization. This

corresponds to the mixing trend between residual Fe as goe-thite and hematite, and Mg-silicates. A minor trend corre-sponds to serpentinized rocks and their altered saprolite prod-ucts. For the fresh and serpentinized rocks, a trend shows anincrease of Si concentration which likely corresponds to theformation of talc-like phases, which are richer in Si than ser-pentine and olivine.Well 2 shows loss of Si andMgwhich canbe considered to indicate laterization. The laterization is

Fig. 4 a Interpolated map of the thickness of the laterite cover from the borehole dataset. b Areas where the borehole density is highest and regularlyspaced. C1 and C2 are two perpendicular cross sections showing the lateral and vertical variability of the thickness of the laterite cover

Miner Deposita

Page 157: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

moderate since the loss of Si andMg does not reach more than20 wt.%. Well 3 is characterized by most samples plotting

toward the Si corner, indicating a significant gain in silica.The Si content is much higher than that of the talc-like phases,which suggests occurrence of quartz, confirmed by petro-graphic observations of quartz in microfractures. It is notewor-thy that in well 3, even the samples rich in iron (more than50% hematite or goethite) are enriched in Si, indicating thatquartz crystallization occurred locally.

Lateral variability of the Ni content

For each borehole, the average element content was calculatedin the laterite (when present) and in the saprolite. The use ofthe average concentration reduces the bias that could arisefrom boreholes of different lengths. In the laterite, the aver-aged Ni content is approximately constant (Table 1). In thesaprolite, the averaged Ni content is more variable, between0.6 and 2.3 wt.% (Table 1).

Figure 8b shows the variability of the averaged percentageby weight of the Ni content in the saprolite at the scale of themining area along the ridge of the massif. The density ofboreholes is highly variable laterally. In areas with a largenumber of boreholes, Ni-rich and Ni-poor areas are organizedas hectometric patches with variable laterite thickness. Thereis an inverse correlation between the thickness of the lateritecover and the Ni content of the underlying saprolite. We di-vided this area in three sub-areas (1, 2, 3; Fig. 8c), where therelationship between the thickness of the laterite cover and theNi content is well expressed. In sub-area 1, the zone with nolaterite cover is slightly enriched in Ni with an average Nicontent between ∼1.5 and 2 wt.%. In sub-area 2, average Nicontent is between ∼2.5 and 4 wt.% in the zonewith no lateritecover, whereas the average Ni content is near 1.5 wt.% underthick laterite cover. Finally, in sub-area 3, the relationshipbetween thickness and Ni content is more complex with areasrich in Ni localized both under thick and thin laterite cover.

Behavior of elements from isocon diagrams

Mass balance calculations were carried out using the isoconmethod of Grant (1986) and Brimhall and Dietrich (1987).This approach has been successfully applied to other Ni later-ite deposits (Brimhall and Dietrich 1987; Gleeson et al. 2004;Fu et al. 2014) and offer insights into the processes associatedwith the laterite formation.

Selection of immobile elements

Several representative samples of laterite, saprolite, orprotolith from boreholes, wells 1, 2 and 3, were investigatedusing the isocon method (Fig. 9; Online Resource 2)

For harzburgite, Ti, Zr, Th, Al, Cr, and Fe are consid-ered immobile elements during weathering (Golightly2010). Ti, Th, and Zr, were not measured, whereas Al

Fig. 5 a Field view of the laterite-saprolite interface. b, c Detailed 3Dmodel of the interface (no vertical exaggeration), showing multiple small-scale depressions, of metric to decametric vertical amplitude

Table 1 First and 3rd quartiles of the bedrock, saprolite, and laterite asdocumented from the whole borehole dataset

Element (wt.%) Bedrock Saprolite Oxide laterite

SiO2 40.3–42 40.5–44.8 2.5–21

Al2O3 0.48–0.62 0.57–1 3.4–5.5

Fe2O3 7.6–8.3 8.6–13.5 53–70

MgO 39.2–42 26–37.3 0.8–3.2

Ni 0.25–0.36 0.6–2.3 1.1–1.7

Number of samples 4260 201,027 42,427

Miner Deposita

Page 158: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

and Cr have low and variable concentrations in thebedrock. Golightly (2010) suggested that even if Fe ismore mobile than Cr, it may be a reliable low-mobilityelement. Indeed, Fe forms secondary minerals during lat-er i te format ion, as demonst ra ted by ubiqui tousFe-oxy-hydroxides. The validity of this assumption may,however, be questioned since low pH and organic matter

are known to increase the transfer of insoluble elements(Ma et al. 2007). For instance, Ma et al. (2007), using Thas an immobile element, documented Fe mobility of 40%in the first meter of a laterite profile developed on basalt,decreasing to around 20% in the following few meters. AtKoniambo, analysis of groundwater (Jeanpert andDewandel 2013) indicates pH of 5.5 near the surface to a

Fig. 6 Variation of major and some trace elements through three typicallaterite profiles with different thicknesses of laterite cover. Their locationsare shown in Fig. 4. Chemical analyses are reported in Online Resource 1.

The black dashed lines indicate the several locations that we consider asrepresentative of laterite, saprolite, and protolith composition (see alsoOnline Resource 2), used for isocon mass balance (Fig. 9)

Fig. 7 Ternary Si-Mg+ Ni-Fediagram of chemical analyses forsamples from the three typicallaterite profiles

Miner Deposita

Page 159: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

value of 10 at depth, either in the laterite or in theserpentinized peridotite. Jeanpert and Dewandel (2013) in-dicate negligible Fe in groundwater. Locally, water with Fecontents greater than 50 mg/L has been measured, but thishigh content likely relates to the presence of small Fe-rich

particles, associated with the dismembering of duricrusts.Similarly, analyses of groundwater from the Massif du Sud(Trescases 1969) indicate that among 15 elements (includ-ing Ti), Fe is the less soluble. In summary, it is likely thatFe may display limited mobility close to the surface.

Fig. 8 a Location of the mining exploited area (pink rectangle) in theKoniambo Massif (represented by the green line). b Average Ni contentin the saprolite calculated for each borehole in the main mining area. Theareas with a predominant blue color have a low Ni content (<2 wt.%),

whereas areas with white to red color have high Ni content (>2 wt.%).The pink rectangle delimits the area presented in c. c Focus on the Ni-richareas where borehole distribution is mostly homogeneous and regularlyspaced. The area of greatest laterite thickness is also represented

Miner Deposita

Page 160: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Moreover, the fact that Fe, Al, and Cr lie along the sameisoline in the 3 wells (Fig. 9) strengthens the interpretationthat Fe is a valid invariant for mass balance computations.

Elements behavior along the 1D vertical profiles

In laterite cover (wells 1 and 3, Fig. 9a, f; Online Resource 2),Mg and Si are almost entirely leached. Ni is depleted with aloss of around 50 to 70% with respect to the parent rock. Forboth wells, the volume ratio between altered and protolith(Eq. 3 in the Online Resource 3) is high, ranging from 6.5 to7.2.

In saprolite (Fig. 9c–e, g), Mg and Si are systematicallydepleted with relative losses for Mg ranging from 25 to 95%

and for Si from 20 to 70% (Online Resource 2). The Ni rela-tive enrichment ranges from 40 to 80% in the BNi-poorsaprolite^ and between 290 and 460% in the BNi-richSaprolite^ (Fig. 9c, d, g; Online Resource 2). The highestenrichment for Ni is in wells 2 and 3 in saprolite covered bythin or without laterite. The volume change associated withsaprolite formation ranged from 1.2 to 3, which is smaller thanfor laterite (Online Resource 2) and consistent with the factthat saprolite is an intermediate state of alteration.

Ni 1D mass balance

We applied a 1D (vertical) Ni mass balance calculation oneach borehole (Fig. 10 and Online Resource 3) by comparing

Fig. 9 Isocon diagrams based on the composition of representativesamples (location in Fig. 4) of laterite and saprolite for the three typicallaterite profiles. The black line represents the isoline drawn to fit leastmobile elements (Fe, Al, Cr). The number indicates the value of the

volume change factor calculated considering Fe as immobile. The blackdashed line represents the constant mass isocon. The chemical composi-tion is from Online Resource 1. The isocon calculations are in OnlineResource 2

Miner Deposita

Page 161: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the Ni mass contained in the weathered column with thatinitially contained in the protolith column deduced from Fecontent.

The variation of the Ni mass balance calculated for each ofthe 6700 boreholes as a function of the thickness of the lateritecover (Fig. 11) shows that the boreholes depleted in Ni arerare, the majority of them being either enriched in Ni (as high

as 1300%). The boreholes with the thickest laterite cover(>25 m) have, at first order, 80–120% Ni enrichment but with40% of boreholes slightly depleted in Ni.

Figure 12b shows a map of the Ni mass balance calcu-lations in the mining area. Two types of areas can be iden-tified organized as hectometric-sized patches. The first hasa Ni mass balance ranging from ∼80 to 120% such that theNi mass can be explained by a per descensum weatheringof the protolith. The second type of area has a Ni massbalance ranging between ∼150% and at least 300%, whichindicates that they are significantly enriched in Ni com-pared to the mass of Ni initially contained in the protolith.It is noteworthy that these patches correlate with the dis-tribution of laterite thickness (Figs. 4a, b and 12c). Indeed,Ni enrichment is typical of thin cover area whereas forthick cover, Ni mass balance is between 80 and 120% ordepleted. In Ni-richest areas, where borehole density ishigh enough and regularly spaced, the relationship be-tween thickness of the laterite cover and Ni mass balanceis well expressed, whereas in areas where the Ni massbalance is 80–120%, the laterite cover is thick (Fig. 12c).Figure 13 shows the distribution of the 80–120% andenriched (>150%) areas compared to the relief. TheNi-balanced areas are localized on topographic highs cor-responding to areas where the laterite is thickest, asattested by duricrusts (Fig. 13a, b). Inversely, theNi-enriched areas, under thin or lack of laterite cover, arelocalized on gentle slopes, downward of the Ni-balancedareas (Fig. 13a, b). Figure 13c shows a slightly differentgeometry where the 80–120% area is localized on a steepslope and downstream of enriched areas. It is important toconsider that a significant bias in the Ni mass balancecalculation remains, which requires to interpret only inrelative enrichment for the areas with thin or lack oflaterite cover. For the thick laterite cover zones on thetopographically high areas, it seems reasonable to

Fig. 10 Schematic sketch of the principle of the 1D vertical, Ni massbalance calculation

Fig. 11 Diagram showing thevariation of the Ni mass balanceas a function of the thickness ofthe laterite cover

Miner Deposita

Page 162: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

consider that the laterite profile is complete or nearlycompletely developed. The Ni mass balance can beconsidered quantitative given that no or little materialwas lost by mechanical erosion. However, for the areas

with thin laterite cover, mainly localized on slopes, thecurrent thickness must be considered as a minimalthickness given that these areas have probably beenaffected by mechanical erosion. The erosion of a part of

Fig. 12 a Location of the mining area (pink rectangle) in the KoniamboMassif (represented by the green line). b Ni mass balance map for eachborehole in the main mining area. The blue color indicates the 80–120%interval whereas the red color indicates areas of significant enrichment.

The pink rectangle delimits the area presented in c. c Area enriched in Niwhere borehole distribution is mostly homogeneous and regularly spaced.The area of thickest cover is also represented

Miner Deposita

Page 163: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

the laterite cover should lead to an overestimation of theenrichment in Ni compared to the protolith. Indeed, theratio of volume decrease induced by the laterization pro-cess is ∼6–7 (Online Resource 2). Consequently, erosion of1 m of laterite implies that the Ni initially in a protolithcolumn of about 6–7 m (Online Resource 2) is not takeninto account in the Ni mass balance calculation.

Discussion

Present day geometry of the laterite cover

The thickness of the laterite cover varies at different scales. Ata large scale, its thickness is a function of the slope of the areawhere it occurs (Figs. 4(c1, c2) and 13). This distribution has

Fig. 13 Distribution of the 80–120% interval and enriched areas determined by Ni mass balance (1.5 vertical exaggeration)

Miner Deposita

Page 164: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

already been described in New Caledonia by several authors(de Chetelat 1947 his Fig. 7; Trescases 1975; Avias 1978;Maurizot et al. 2002; Beauvais et al. 2007). The thickest coveroccurs mainly on topographic highs and consists of a com-plete lateritic profile preserved from erosion. On gentle slopes,the laterite cover has moderate thickness. These profiles areincomplete which is probably the result of erosion (Latham1975; Trescases 1975; Avias 1978; Maurizot et al. 2002).Consequently, it is difficult to estimate the initial thicknessof the laterite cover localized on slopes. Moreover, these lat-erites are also, at least in part, composed of reworkedmaterials(Latham 1975; Trescases 1975; Avias 1978; Maurizot et al.2002). This allochthonous material may have been derivedfrom the erosion of ancient or of present-day thick lateriteslocalized upstream. For the Koniambo Massif, thick profileslocalized on topographic highs can be considered as nearlycomplete because the duricrust is only slightly removed(Maurizot et al. 2002), and thus, the erosion of ancient lateriteprofiles seems most likely. This is consistent with the origin ofreworked materials proposed by Beauvais et al. (2007) forlaterites on slopes of the Tiébaghi Massif (Fig. 2). Therefore,the present-day geometry of the laterite shows two end-mem-bers: (1) complete, thick, laterite, mainly localized on topo-graphic highs and (2) thin laterite, localized downstream of thetopographic highs, which have been partially eroded andreworked.

At a smaller scale, the interface between the lateritecover and the saprolite is irregular with local depressionsof metric to decametric vertical amplitude (Figs. 4(c1) and5a, b). This geometry could reflect variations in density offractures affecting the bedrock. Indeed, intense fracturingof the upper part of the Peridotite Nappe is considered oneof the main parameters which led to weathering of thebedrock (Trescases 1975; Leguéré 1976). Consequently,areas with a high density of fractures are most likely todevelop an efficient drainage system leading to develop-ment of (i) preferential dissolution zones leading tokarstification and therefore (ii) local deepening of the lat-erite profile. At a larger scale, the laterite-saprolite inter-face can form endorheic areas (Fig. 5c) which can beinterpreted as dolines where material could be, in part,autochthonous but also composed of reworked materialsderived from erosion of ancient laterites localized in thevicinity (Latham 1975; Trescases 1975; Genna et al. 2005).

The Tiébaghi Massif, north of Koniambo (Fig. 2), is knownto have a massive and continuous lateritic cover on the plateau(Beauvais et al. 2007) whereas the Koniambo laterite cover ishighly dissected by erosion, forming high relief morphology.The massive laterite cover is also affected by erosion which ismainly localized at the plateau margins (Beauvais et al. 2007).Our 3D model supports the hypothesis that the thick lateriteareas can be extrapolated from one area to the neighboringone. This suggests that these surfaces represent relicts of a

past, massive, and continuous laterite cover as the one ob-served in the Tiébaghi massif.

From the Fe low mobility and the Ni 1D mass balancecalculations, it is possible to estimate the paleothickness ofthe peridotite responsible for the observed amount of Ni ina laterite. If these estimates are limited to the wells whereNi is vertically balanced, the computed thicknesses arevariable, ranging from 0 to 300 m, with an average valueof 100 m. For example, paleothickness of 100 and 300 mwould correspond to a profile with 11 m of laterite above21 m of saprolite or 40 m of laterite above 50 m of sapro-lite, respectively. However, the proportions of laterite andsaprolite vary and the highest values are found alongpresent-day ridges and are associated with the thickestand best preserved duricrust/laterite layers. On the otherhand, it is also interesting to compare these values withweathering rates (29 mm per 1000 years) calculated byTrescases (1975) in New Caledonia for plateaus.Therefore, this rate would require ∼10 Myr to weather athickness of 300 m or ∼3.5 Myr for the average value of100 m. When considering these values, two points shouldbe taken into account. First, the estimated paleothickness ismoderate compared to the present-day thickness of the pe-ridotite of the Koniambo Massif (ca. 800 m). Second, theperidotite obduction ended before ∼27 Ma (Cluzel et al.2012), such that 3.5 to 10 Myr is a relatively short periodof time. Thereafter, based on the present-day geometry,these values probably document a single and relativelyshort episode of weathering affecting a moderate volumeof peridotite. The lateritic surfaces at various elevations onthe present-day relief, in the Koniambo Massif and else-where on the island, probably developed during otherweathering and/or tectonic episodes (Chevilotte et al.2006; Sevin et al. 2012).

Significance of Ni variability

In the laterite zone, the Ni concentration is low constant exceptabove the saprolite zone, where Ni concentration can slightlyincrease. The main part of the Ni mineralization is in thesaprolite zone where the average Ni concentration is∼2 wt.%. The Ni concentration is variable vertically acrossthis zone (Fig. 6), which reflects the complexity of the pro-cesses involved in Ni redistribution. Laterally, Ni grade in thesaprolite zone is also variable both at large (×100m) and small(×1 m) scales (Fig. 8).

At the small scale, Ni-rich and Ni-poor areas form metricpatches without clear organization. Several parameters are in-voked in literature to explain such variations (Butt and Cluzel2013). In our case, heterogeneity in the bedrock compositioncannot be invoked because Ni variance in bedrocks is low. Bycontrast, the structure of the bedrock can influence the local-ization of the Ni mineralization. Areas of high fracture density

Miner Deposita

Page 165: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

likely provided local, preferential drainage leading to localdeepening of weathering and increase in the Ni content.Some faults can form barriers to water flow leading to prefer-ential concentration of Ni along them (Leguéré 1976). On theother hand, Cathelineau et al. (2016a) showed that recent topresent-day lateral redistribution of nickel can also occur inthe saprolite zone. This lateral redistribution is the result ofwater table fluctuations leading to dissolution of early garni-erite and to formation of thin layers of silicates cementingjoints, the so-called target-like ore. At a large scale, Ni-richand Ni-poor areas form hectometric-sized patches whichbroadly correlate with the present-day laterite thickness(Fig. 8). The richest areas are below thin or lack of lateritecover whereas the poorest are under the thickest laterite cover.This is consistent with earlier observations by de Chetelat(1947) and Avias (1969), who identified Ni-rich areas underthin or absent laterite cover on slopes. This has also beenreported by Gleeson et al. (2004) for the Ni laterite depositof Cerro Matoso in Colombia. This result seems paradoxicalgiven that the main part of the Ni mineralization is thought toresult from downward mobility of Ni (Glasser 1904; deChetelat 1947; Avias 1969; Trescases 1975). In this model,Ni leached from the weathered protolith is progressively in-corporated in Ni-bearing phases to enrich the saprolite zone.Following this model, a higher enrichment in Ni in saprolite isexpected beneath the thickest laterite cover.

We suggest that lateral redistribution of nickel could ex-plain, at least in part, the geometry of the laterite nickel depositof the Koniambo Massif. On the one hand, mechanical trans-port of laterite materials on slopes, including Ni-bearing de-bris, could explain local enrichments in the saprolite zone.Evidence of reworked material was identified in thin lateritecover on slopes (Latham 1975; Trescases 1975; Avias 1978;Maurizot et al. 2002). Even if erosion is active on slopes,successive leaching of laterite materials, coming from formerlateritic profiles, could lead to a significant increase in Nigrade in the underlying saprolite zone (Fig. 14a). However,given that the past relief of the massif and the flux of mechan-ical erosion of lateritic material l are unknown, it is difficult toestimate the nickel input linked to this process. On the otherhand, lateral infiltration of groundwater enriched in dissolvedNi from thick laterite profiles in topographic highs could alsoexplain local enrichments in the saprolite zone downstream(Fig. 14b). This process has already been invoked in NewCaledonia based on the fact that the different zones of thelateritic profile, which have different hydraulic transmissivity,can cause groundwater infiltration (de Chetelat 1947; Avias1969; Trescases 1975; Join et al. 2005). At a first order, the Nimass balance study at the deposit scale shows that areas withcomplete lateritic profiles have mass balance mainly between80 and 120% (Fig. 12c). This suggests that these areas did notundergo significant loss of nickel. However, Fig. 11 highlightsthat 40% of boreholes are slightly depleted in nickel where

laterite cover is thicker than 25 m. Integrated at the scale of athick laterite cover area, the sum of these losses could repre-sent a significant mass of nickel. This supports lateral infiltra-tion of groundwater rich in dissolved Ni as a process leadingto the significant enrichment of saprolite in slope areas. Only afew data on Ni content in groundwater are available in litera-ture. In the Dumbéa Basin (Fig. 2), the Ni content is in therange 4–10 μg L−1 (Trescases 1969) whereas the PiroguesRiver basin (Fig. 2) is in the range 10–30 μg L−1 (Peiffertet al. 2004). These values indicate that a part of the dissolvednickel is not trapped by oxy-hydroxide or Ni silicates and canbe laterally transported by water in solution.

Conclusion

On the Koniambo Massif, two main types of lateritic pro-files are recognized. The first type is thick and nearlycomplete laterite sections, localized on topographic highs.The second type is thin and incomplete laterite profileslocalized on gentle slopes (<20°). In contrast to the

Fig. 14 Sketches illustrating the two possibilities invoked in the text toexplain Ni enrichment in the saprolite from the slope. a Successiveleaching of laterite materials, coming from former lateritic profiles,leading to an increase in Ni grade in the underlying saprolite. b Lateralinfiltration of groundwater enriched in dissolved Ni from thick lateriteprofiles in topographic highs to the saprolite zone downstream

Miner Deposita

Page 166: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

classical per descensum model of Ni laterite deposit for-mation, we show that areas richest in Ni are under thin orlack of laterite cover. Based on the Ni mass balance foreach borehole on a kilometric scale, we propose that pro-cesses of lateral nickel transfer lead to an increase in Nigrade in the saprolite, on gentle slope areas downstreamof ridges. At least two complementary processes can beinvoked. The first consists in the successive lateral trans-port of reworked laterite material, including Ni-bearingmaterial, from topographic highs to downstream slopeareas. Leaching of this reworked material could increaseNi grades in the underlying saprolite zone. The secondconsists lateral infiltration of groundwater rich in dis-solved Ni from high topographic areas to downstreamslope areas.

Acknowledgements This study constitutes a part of B. Quesnel PhDwhich was supported by Koniambo S.A.S. GeoRessources contributionhas been supported by ANR-10-LABX-21-LABEX RESSOURCES21.We gratefully thank Martin A. Wells and an anonymous reviewer fortheir detailed and constructive reviews and Beate Orberger and GeorgesBeaudoin for their editorial handling and their constructive comments.

References

Audet M-A (2008) Le massif du Koniambo, Nouvelle-CalédonieFormation et obduction d’un complexe ophiolitique du type SSZ.Enrichissement en nickel, cobalt et scandium dans les profilsrésiduels. Thèse, Université de la Nouvelle-Calédonie etUniversité du0 Québec à Montréal: pp. 355.

Avias J (1967) Overthrust structure of the main ultrabasic NewCaledonian massives. Tectonophysics 4:531–541

Avias J (1969) Note sur les facteurs contrôlant la genèse et la destructiondes gîtes de nickel de la Nouvelle-Calédonie. Importance desfacteurs hydrologiques et hydrogéologiques C R Acad Sc Paris268:244–246

Avias J (1978) L’évolution des idées et des connaissances sur la genèse etsur la nature des minerais de nickel, en particulier latéritiques, deleur découverte à nos jours. Bulletin B.R.G.M (Bureau desRecherches Géologiques et Minières), section II (3) :165–172.

Baille J-S, Jegat P, Marini D, Louis A, (2013) Atlas de la Nouvelle-Calédonie, IRD éditions, 269 pp. N°EAN : 9782709917407

Beauvais A, Parisot J-C, Savin C (2007) Ultramafic rock weathering andslope erosion processes in a South West Pacific tropical environ-ment. Geomorphology 83:1–13

Brimhall GH, Dietrich WE (1987) Constitutive mass balance relationsbetween chemical composition, volume, density, porosity, and strainin metasomatic hydrochemical systems: results on weathering andpedogenesis. Geochim Cosmochim Acta 51(3):567–587

Butt CRM, Cluzel D (2013) Nickel laterite ore deposits: weatheredserpentinites. Elements 9(2):123–128

CathelineauM,Quesnel B, Gautier P, Boulvais P, Couteau C, Drouillet M(2016a) Nickel dispersion and enrichment at the bottom of the reg-olith: formation of pimelite target-like ores in rock block joints(Koniambo Ni deposit, New Caledonia). Mineral Deposita 51:271–282

Cathelineau M, Myagkiy A, Quesnel B, Boiron M-C, Gautier P, BoulvaisP, Ulrich M, Truche L, Golfier F, Drouillet M (2016b) Multistagecrack seal vein and hydrothermal Ni enrichment in serpentinized

ultrmafic rocks (Koniambo massif, New Caledonia). Accepted withminor revisions in mineralium deposita.

Chevillotte V, Chardon D, Beauvais A, Maurizot P, Colin F (2006) Long-term tropical morphogenesis of New Caledonia (Southwest Pacific):importance of positive epeirogeny and climate change.Geomorphology 81(3–4):361–375

Cluzel D, Vigier B (2008) Syntectonic mobility of supergene nickel oresof New Caledonia (Southwest Pacific). Evidence from garnieriteveins and faulted regolith. Resour Geol 58:161–170

Cluzel D, Aitchison J-C, Picard C (2001) Tectonic accretion and under-plating of mafic terranes in the Late Eocene intraoceanic fore-arc ofNew Caledonia (Southwest Pacific): geodynamic implications.Tectonophysics 340:23–59

Cluzel D, Maurizot P, Collot J, Sevin B (2012) An outline of the geologyof New Caledonia; from Permian-Mesozoic SoutheastGondwanaland active margin to Cenozoic obduction and supergeneevolution. Episodes 35:72–86

de Chetelat E (1947) La genèse et l’évolution des gisements de nickel dela Nouvelle-Calédonie. B. Soc. Geol. Fr. 5 (XVII):105–160.

Dublet G, Juillot F, Morin G, Fritsch E, Fandeur D, Ona-Nguema G,Brown GE (2012) Ni speciation in a New Caledonian lateritic reg-olith: a quantitative X-ray absorption spectroscopy investigation.Geochim Cosmochim Acta 95:119–133

Dublet G, Juillot F, Morin G, Fritsch E, Fandeur D, Brown GE (2015)Goethite aging explains Ni depletion in upper units of ultramaficlateritic ores from New Caledonia. Geochim Cosmochim Acta160:1–1

Freyssinet P, Butt CRM, Morris RC, & Piantone P (2005) Ore-formingprocesses related to lateritic weathering. Society of EconomicGeologists, Economic Geology 100th Anniversary Volume:681–722.

Fritsch EM, Juillot FA, Dublet GA, Fonteneau L, Fandeur D, Martin E,Caner L, Auzende A-L, GraubyO, Beaufort D (2016) An alternativemodel for the formation of hydrous Mg/Ni layer silicates(Bdeweylite^/Bgarnierite^) in faulted peridotites of New Caledonia:I. Texture and mineralogy of a paragenetic succession of silicateinfillings. Eur. J. Mineral 28(2):295–311. doi:10.1127/ejm/2015/0027-2503

Fu W, Yang J, Yang M, Pang B, Liu X, Niu H, Huang X (2014)Mineralogical and geochemical characteristics of a serpentinite-derived laterite profile from East Sulawesi, Indonesia: implicationsfor the lateritization process and Ni supergene enrichment in thetropical rainforest. J Asian Earth Sci 93:74–88

Genna A, Maurizot P, Lafoy Y, Augé T (2005) Role of karst in thenickeliferous mineralisations of New Caledonia. C R Geoscience337:367–374

Glasser E (1904) Rapport à M le ministre des Colonies sur les richessesminérales de la Nouvelle-Calédonie. Paris, Annales des Mines,560p.

Gleeson SA, Herrington RJ, Durango J, Velasquez CA, Koll G (2004)The mineralogy and geochemistry of the Cerro Matoso S.A. Nilaterite deposit, Montelibano, Colombia. Econ Geol 99(6):1197–1213

Golightly JP (1981) Nickeliferous laterite deposits. Economic Geology,75th Anniversary Volume: 710–735.

Golightly JP (2010) Progress in understanding the evolution of nickellaterites. Society of economic Geologists, Special Publications 15

Grant JA (1986) The isocon diagram—a simple solution to Gresens’equation for metasomatic alteration. Econ Geol 81:1976–1982

Guillon JH (1975) Les massifs péridotitiques de Nouvelle-Calédonie:type d'appareil ultrabasique stratiforme de chaine récente.Mémoire ORSTOM (Office de la Recherche Scientifique etTechnique Outre-Mer) 76:120

Jeanpert J, Dewandel B (2013) Analyse préliminaire des donnéeshydrogéologiques du massif du Koniambo. Public ReportBRGM/RP-61765-FR, 102 pp.

Miner Deposita

Page 167: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Join JL, Robineau B, Ambrosi JP, Costis C, Colin F (2005) Systèmehydrogéologique d’un massif minier ultrabasique de Nouvelle-Calédonie. C R Géosci 337(16):1500–1508

Latham M (1975) Géomorphologie d’un massif de roches ultrabasiquesde la côte Ouest de la Nouvelle Calédonie : le Boulinda. Cah.ORSTOM. Sér. Géol. VII(1):17–37

Latham M (1977) On geomorphology of northern and western new cal-edonian ultramafic massifs. In: International symposium ofgeodynamics in South West Pacific. Nouméa, Technip Paris éd, pp235–244

Latham M (1986) Altération et pédogenèse sur roches ultrabasiques enNouvelle-Calédonie. PhD thesis, Univ. Dijon. Dijon, France. 331 p.

Leguéré J (1976) Des corrélations entre la tectonique cassante etl’altération supergène des péridotites de Nouvelle Calédonie. PhDthesis, Univ. Du Languedoc, Montpellier, France, 95p.

Ma JL, Wei GJ, Xu YG, LongWG, SunWD (2007)Mobilization and re-distribution of major and trace elements during extreme weatheringof basalt in Hainan Island, South China. Geochimica CosmochimicaActa 71(13):3223–3237

Mallet J-L (2002) {G}eomodeling. Oxford University Press, New York,NY, Applied Geostatistics.

Maurizot P, Lafoy Y, Poupée M (2002) Cartographie des formationssuperficielles et des aléas mouvements de terrain en Nouvelle-Calédonie, Zone du Koniambo. Bureau de RecherchesGéologiques et Minières, Public Report RP51624-FR, 45 p.

Maurizot P, Vendée-Leclerc M (2009) Carte géologique de la Nouvelle-Calédonie au 1/500.000. Service Géologique de Nouvelle-Calédonie/DIMENC-BRGM, Nouméa

Orloff O (1968) Etude géologique et géomorphologique des massifsd’ultrabasites compris entre Houailou et Canala (Nouvelle-Calédonie). Ph.D. thesis, Université de Montpellier, France, 189 pp.

Paris J-P (1981) Géologie de la Nouvelle-Calédonie, un essai de synthèse.Mémoires du Bureau de Recherches Géologiques et Minières 113:274

Pelletier B (1989) Les minerais de nickel de Nouvelle-Calédonie. SociétéLe Nickel, 10 pp.

Peiffert C, Bouichet G, Traore D, Cathelineau M, Colin F (2004) Ni, Cr,and Co transfer during the erosion of laterites developed on ultraba-sic formations (Rivière des Pirogues area, New Caledonia). RST 20-25 septembre 2004, Strasbourg, RSTGV-A-00405

Prinzhofer A, Nicolas A, Cassard D, Moutte J, Leblanc M, Paris JP,Rabinovitch M (1980) Structures in the New Caledonia peridotites-gabbros: implications for oceanic mantle and crust. Tectonophysics69:85–112

Quesnel B, Gautier P, Boulvais P, Cathelineau M, Maurizot P, Cluzel D,Ulrich M, Guillot S, Lesimple S, Couteau C (2013) Syn-tectonic,meteoric water-derived carbonation of the New Caledonia peridotitenappe. Geology 41(10):1063–1066

Quesnel B, Boulvais P, Gautier P, Cathelineau M, John CM, Dierick M,Agrinier P, Drouillet M (2016a) Paired stable isotope (O, C) andclumped isotope thermometry of magnesite and silica veins in theNew Caledonia Peridotite Nappe. Geochim Cosmochim Acta 183:234–249

Quesnel B, Gautier P, Cathelineau M, Boulvais P, Drouillet M (2016b)The internal deformation of the Peridotite Nappe of New Caledonia:a structural study of serpentine-bearing faults and shear zones in theKoniambo Massif. J Struct Geol 85:51–67

Schellmann W (1989) Composition and origin of lateritic nickel ore atTagaung Taung, Burma. Mineral Deposita 24:161–168

Sevin B, Ricordel-PrognonC, Quesnel F, Cluzel D, Lesimple S,MaurizotP (2012) First palaeomagnetic dating of ferricrete in NewCaledonia:new insight on themorphogenesis and palaeoweathering of BGrandeTerre^. Terra Nov. 24(1):77–85

Soler JM, Cama J, Gali S, Meléndez W, Ramirez A, Estanga J (2008)Composition and dissolution kinetics of garnierite from the Loma deHierro Ni-laterite deposit, Venezuela. Chem Geol 249:191–202

Trescase J-J (1969) Géochimie des altérations et des eaux de surface dansle Massif ultrabasique du Sud de la Nouvelle-Calédonie. Bull ServCarte Géol Als Lorr 23(4):329–354

Trescases J-J (1975) L’évolution géochimique supergène des rochesultrabasiques en zone tropicale : formation des gisementsnickélifères de Nouvelle-Calédonie. Mémoires ORSTOM (Officede la Recherche Scientifique et Technique Outre-Mer) 78:259

Ulrich M, Picard C, Guillot S, Chauvel C, Cluzel D, Meffre S (2010)Multiple melting stages and refertilization as indicators for ridge tosubduction formation : the New Caledonia ophiolite. Lithos 115:223–236

Ulrich M (2010) Péridotites et serpentinites du complexe ophiolitique dela Nouvelle-Calédonie. Ph.D thesis, Université de Nouvelle-Calédonie et Université Joseph Fourier (Grenoble), France, 246 pp.

USGS (U.S Geological Survey) (2016) Mineral commoditiy sum-m a r i e s , J a n u a r y 2 0 1 6 . h t t p : / / m i n e r a l s . u s g s .gov/minerals/pubs/commodity/nickel/mcs-2016-nicke.pdf

Villanova-de-Benavent C, Proenza JA, Galí S, García-Casco A, Tauler E,Lewis JF, Longo F (2014) Garnierites and garnierites: textures, min-eralogy and geochemistry of garnierites in the Falcondo Ni-lateritedeposit, Dominican Republic. Ore Geol Rev 58:91–109

Xstrata (2012) Mineral resources and reserves at 31 Dec. 2012. AnnualReport, 50 p

Miner Deposita

Page 168: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite
Page 169: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Genesis and mineralogy of Ni ores

Myagkiy A., Truche, T. , Cathelineau M., Golfier F. (2017)

Revealing the conditions of Ni mineralization in the laterite profiles of New Caledonia: Insights from reactive geochemical

transport modelling..

- 12 -

Chem. Geol.466, 274-284.

Page 170: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Contents lists available at ScienceDirect

Chemical Geology

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

Revealing the conditions of Ni mineralization in the laterite profiles of NewCaledonia: Insights from reactive geochemical transport modelling

Andrey Myagkiya,*, Laurent Truchea,b, Michel Cathelineaua, Fabrice Golfiera

a GeoRessources Lab., UMR 7359 of CNRS, CREGU, University of Lorraine, F-54518 Vandoeuvre-Lès-Nancy Cedex, Franceb ISTerre, UMR 5275 of CNRS, University of Grenoble Alpes, Grenoble Cedex 9 F-38041, France

A R T I C L E I N F O

Keywords:LateriteTransport modellingWeatheringKineticsWater-rock interactions

A B S T R A C T

New Caledonia is one of the world's largest nickel laterite deposits that form from intense chemical and me-chanical weathering of a peridotite bedrock. As a result of such a weathering process a subsequent downwardmigration of Si, Mg and Ni takes place, which eventually leads to redistribution of the elements in depth and overtime depending on their mobility. Being released from ultramafic parent rock to groundwater, the mobility ofnickel is to a great extent controlled by sorption, substitution and dissolution/precipitation processes. Therefore,the final profile of nickel enrichment is the result of the superposition of these possible fates of nickel. The wayhow Ni is redistributed in between them represents and defines its mineralization in laterite. Knowledge of theseprocesses along with factors, controlling them appears to be a key to detailed understanding of laterite for-mation. In this study a numerical model, which solves the reaction-transport differential equations, is used tosimulate the formation of laterite profile from ultramafic bedrock with particular emphasis on modelled Nienrichment curve, its comparison with in situ observations, and detailed understanding of trace elements mo-bility. Since nickel deposits in New Caledonia is characterized by oxide and hydrous Mg silicate ores, threedifferent concurrent fates of Ni deposition in a profile were taken into account in the modelling: i) Ni in agoethite crystal lattice, ii) Ni sorbed on weak and strong goethite sorption sites, and iii) Ni precipitated withsilicates (garnierite). Simulations were performed using PHREEQC associated with llnl.dat thermodynamic da-tabase that has been edited in order to account garnierite minerals used in the calculations. The work outline isrepresented by: i) long term (10 Ma) simulation of nickel laterite formation and evolution, ii) analysis of mobilityof the elements and understanding its controlling factors, iii) comparison of modelled and in situ Ni enrichmentprofile and analysis of nickel distribution in between different retention processes, iv) modelling and in depthunderstanding of these retention processes. The modelling reveals that the vertical progression of the pH frontcontrols thickening of iron-rich zone, explains the vertical mobility of the elements and governs the Ni en-richment. Adsorption itself plays an important role in lateritization process retarding Ni mobility, but i) becomessignificant in a narrow range of pH (slightly alkaline) due to competition of Mg and Ni for sorption sites and ii)does not explain such a high nickel content in limonite nowadays, suggesting that Ni is held in goethite mostly bystronger ties i.e. substituted for Fe in the crystal lattice of iron oxide. 1-D modelling appears to be a powerful toolin understanding the general behavior of trace elements upon the formation of laterite and at the same timereveals that locally Ni mineralizations should be explained by more complex processes, such as lateral transfers,convective flows and preferential pathways.

1. Introduction

Nickel laterite ores account for over 60% of global Ni supply (Buttand Cluzel, 2013), thus, making it an important exploration target inanticipation of the future demand of Ni. One of the biggest nickeliferouslaterite reserves in the world is located in New Caledonia, 1300 km eastof Australia in the southwest Pacific Ocean. Representing around 30%of the global resources in nickel, these laterites are actively worked by

various mining companies (SLN-ERAMET, KNS, SMSP etc.). They occurat the top of the obducted ultramafic massif with an average thicknessof 30–40 m, resulting from the weathering of the Peridotite Nappe. Theinitial nickel content of the parent rock, that consists of olivine andvarious form of serpentine (antigorite, chrysotile, lizardite, polygonalserpentine), can contain up to 0.4 wt% of Ni (Ulrich, 2010; Herzberget al., 2013; Golightly, 1981). The weathering of the olivine-rich ul-tramafic rocks and their serpentinized equivalents by meteoric water

http://dx.doi.org/10.1016/j.chemgeo.2017.06.018Received 28 September 2016; Received in revised form 18 May 2017; Accepted 14 June 2017

* Corresponding author at: ENSG (bat. E), Rue du Doyen Marcel Roubault, TSA 70605, Vandoeuvre-Lès-Nancy Cedex 54518, France.E-mail address: [email protected] (A. Myagkiy).

Chemical Geology 466 (2017) 274–284

Available online 21 June 20170009-2541/ © 2017 Elsevier B.V. All rights reserved.

MARK

Page 171: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

induces the selective leaching of Fe, Mg, Si and Ni and leads to theformation of a lateritic profile together with an exceptional Ni enrich-ment (Trescases, 1975; Golightly, 2010). The limonite level, formed atthe top of the profile (Fig. 1), is mainly composed of iron oxi-hydroxidesand followed by the saprolitic level where hydrous Mg silicate depositsform, especially in the mid to lower saprolite. Ni concentration in-creases constantly throughout the limonite horizon, but the highest Nigrade (locally 2 wt% up to 5 wt%) is found in the saprolite horizonwhich represents about one third of the total Ni laterite resources(Fig. 1). In this horizon, Ni is concentrated in nickeloan varieties ofserpentine, talc-like, chlorite and sepiolite, some of which are poorlydefined and known informally as “garnierite” (Butt and Cluzel, 2013).Transition between the saprolite and limonite horizons is marked by asharp increase in MgO and SiO2 content and, consequently, by drasticchange in mineralogy (Fig. 1). The so-called garnierite is mostly locatedalong fractures (Trescases, 1975). Under the saprolite level, a poorlyaltered peridotite is situated. The basis of the peridotite nappe hostsnumerous pure magnesite veins, sometimes associated with amorphoussilica (Quesnel et al., 2015).

The enrichment of nickel in the weathering profile is governed bymany interplaying factors that include mineralogical composition ofparent rock, climate, chemistry and rates of chemical weathering, tec-tonics, and drainage (Lelong et al., 1976; Golightly, 1981; 2010;Gleeson et al., 2003; Freyssinet et al., 2005; Butt and Cluzel, 2013).After being leached from rock-forming minerals nickel is thought to beredeposited below, mostly associated with i) Ni-rich ferruginous oxideore where it is sorbed and/or substituted for Fe in the crystal lattice(Cornell, 1991; Fischer et al., 2007; Singh et al., 2002; Carvalho-e-Silvaet al., 2003; Dublet et al., 2012), ii) a wide variety of neo-formed sili-cate minerals such as talc-like and sepiolite-like minerals (Cathelineauet al., 2015; Villanova-de-Benavent et al., 2014; Wells et al., 2009), andiii) partly weathered primary lizardite, in which Mg to different extenthas been exchanged by Ni (Golightly, 1981, 2010). Three differentmechanisms have been suggested in the literature in order to explainthe association between nickel and goethite: i) isomorphous substitu-tion of Ni for Fe into the goethite structure, ii) association of Ni withpoorly crystalline phase, and iii) adsorption of Ni to goethite surface.Singh et al. (2002) and Carvalho-e-Silva et al. (2003) found and con-firmed using EXAFS that Ni2+ is associated with goethite via iso-morphous substitution for Fe3+. Using sequential leaching studies ofIndian laterite samples Swamy et al. (2003) reported that the quantityof chemisorbed Ni was negligible and most of the nickel was bound inthe crystal lattice with some associated with amorphous goethite. Ac-tually such an association is quite complex and literature suggests thatthere is a continuum going from one stage to other (Davis et al., 1987;

Stipp et al., 1992). Therefore, an initial fast adsorption step is followedby a slow step of incorporation of the adsorbed species into the crystalstructure with formation of solid solution. Schultz et al. (1987) ex-perimentally documented the irreversible adsorption of metals on fer-rihydrite surfaces. Adsorption/desorption of 10−5 M Ni2+ on 10−3 Mferrihydrite revealed that after the overnight exposure only about 25%of the Ni2+ could be desorbed and the remainder stayed bound in theferrihydrite structure. Similar hysteresis were found by McKenzie(1980), Brümmer et al. (1988), Barrow et al. (1989), and Gerth et al.(1993). Further recrystallization of the ferrihydrite leads to a release ofthe nickel over time. Direct evidence was established for a negativecorrelation between the bulk Ni content and the crystallinity of goethiteas a function of depth in the New Caledonian Ni laterites (Dublet et al.,2012; 2015).

There are a number of scientific articles dealing with modelling ofregolith formation from bedrock. For example Soler and Lasaga (1996,1998) have been working on an advection-dispersion-reaction model ofbauxite formation, Fletcher et al. (2006) and Lebedeva et al. (2007) ongranite bedrock (albite + FeO-bearing phase + quartz) transformationto saprolite (goethite + kaolinite + quartz), Navarre-Sitchler et al.(2011) on a reactive transport model for weathering rind formation onbasalt etc. Nevertheless, the mobility and transport of trace elements inultramafic bedrock was not yet modelled with the only exception of arecent study by Domènech et al. (2017). Their work provides valuableinsights into the formation of the different laterite horizons in theprofile from partially serpentinized peridotite. However, the analysisseems to miss important elements that suggest Ni should be includedinto the goethite structure and cannot be explained by adsorptionalone. In addition, the formation of garnierite is not considered and Nienrichment is shown only in oxide part of deposits. Thus, the processesof Ni retention in a profile and their controlling parameters are stillpoorly understood.

In the present work, we used a reactive transport modelling in orderto study the development of the secondary nickel ores upon verticalprogression of the alteration front. We analyse the effect of key para-meters governing i) sorption processes on a goethite surface, ii) pre-cipitation/dissolution of secondary minerals, and iii) kinetic dissolutionof serpentinized olivine due to meteoric water flow (Fig. 2). Moreover,the effect of pimelite/kerolite (Ni-bearing phases) being present as asolid solution and incorporation of Ni into the goethite structure werealso studied. The role of sorption processes on goethite in Ni redis-tribution in space during the mineralization, as well as the detailedunderstanding of their mechanisms have particular emphasis in thispaper. Direct comparison of Ni distribution obtained after the model-ling with its typical distribution in the profiles provides us in depth

Fig. 1. A schematic log of the New Caledonian PeridotiteNappe correlated with approximate elemental compositionof different zones. Turquoise line represents nickel dis-tribution in a profile. (For interpretation of the referencesto color in this figure legend, the reader is referred to theweb version of this article.)Source: Modified after Ulrich et al. (2011), Troly et al.(1979), and Guilbert and Park (1986).

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

275

Page 172: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

understanding of the weathering process in laterite deposits, reveals thevalidity of the model and gives new insights into the other processesthat might have also governed the mobility of elements and formationof these increasingly important Ni deposits.

2. Materials and methods

2.1. Modelling the hydrodynamic system

A reactive multicomponent 1-D transport model of supergene en-richment of lateritic Ni deposits has been simulated by assuming theweathering of a one-dimensional vertically oriented column of ser-pentinized olivine due to steady state meteoric water flow. The codeused for the simulations is PHREEQC (V 3.1.4) (Parkhurst and Appelo,2013). The 1-D column is defined by a series of 40 cells of 0.5 m inlength. The velocity of water in each cell is determined by the length ofthe cell, porosity and amount of annual rain precipitation. Solute con-centrations at some point on a flowline may change by i) advection ofconcentration gradients, ii) reactions with the solid material, and iii)dispersion and diffusion. The Advection-Reaction-Dispersion equationthat describes these changes along the flowline and implemented onPhreeqc is the following:

⎜ ⎟⎛⎝∂∂⎞⎠

= − ⎛⎝∂∂⎞⎠− ⎛⎝∂∂⎞⎠

+ ⎛⎝∂∂

⎞⎠

ct

ν ct

qt

D cxx t x

Lt

2

2 (1)

where c is the solute concentration (mol⋅l−1), v is pore water flowvelocity (m⋅s−1), DL is hydrodynamic dispersion coefficient (m2 ⋅s−1)and q is the concentration in the solid (mol⋅l−1 of pore water). In ourcalculations we neglect dispersion and set DL to 0. According to in situobservations, indeed, transport of the chemical components is con-trolled by advection and high reaction rates. Neglecting dispersion willcause sharper reaction front and higher concentration peaks. In otherwords, spreading of the concentration front should be slightly under-estimated.

Flux boundary conditions (Cauchy) were defined for the first andlast cell. A slightly acidic tropical rainwater with pH = 5.6 due to itsequilibrium with atmospheric CO2 represents an incoming solution andis injected in the first cell (Fig. 3). Regarding separately the values ofthe experimental rates of kinetic dissolution of the parent rock con-stituents (olivine, serpentine, enstatite), presented in the literature (e.g.Brantley and Chen, 1995; Pokrovsky and Schott, 2000; Wilson, 2004;Thom et al., 2013), one may note that olivine has a highest rate of thedissolution process among them. The rates of enstatite and serpentineare extremely sluggish compared to those of olivine, being one to a few

orders of magnitude lower depending on pH. This makes olivine themain initial mineral in a system that provides nickel and other elements(Mg, Si and Fe). In this way, the numerical modelling was performedassuming olivine in each cell of the 1-D column at initial state (Fig. 2).

The New Caledonia is situated in seasonally humid wet savannas,thus, characterized by summer rainfall of 900–1800 mm and a2–5 months winter dry season (Butt and Cluzel, 2013). Nevertheless,these values are related to the current climatic circumstances and mostlikely were not the same during all 10 Ma of laterite formation. Ac-cording to calculations of Thorne et al. (2012) Ni laterites developwhere rainfall exceeds 1000 mm/y and mean monthly temperaturesrange between 22–31 °C (summer) and 15–27 °C (winter). For our si-mulations the value of 2000 mm/y was chosen as acceptable.

Porosity of the ultramafic bedrock was chosen to be about 2% but ishighly dependent on the fractures density (Join et al., 2005; Jeanpertand Dewandel, 2013). Due to code limitations, the porosity remainsconstant throughout the calculations. There are some other codes withimproved hydrodynamic part (PhreeqcRM, Crunchflow, etc.) thathandle porosity changes but from a 1-D modelling point of view such anassumption has a minor influence on the results of simulation. Indeed,when magnesium and silica are leached from limonite zone a total mass

Fig. 2. Interaction between a rock column and meteoricwater: main variables taken into account.

Fig. 3. Bar chart of chemical composition of incoming solution (rainwater) at equilibriumwith atmospheric CO2 and subsequent pH of 5.6.Source: In accordance with Trescases (1975).

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

276

Page 173: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

loss of up to 70–80% appears there. At the same time, the density aheadof the pH front, where silicates form, will still be much closer to theinitial value than in the goethite zone. Since 1-D column does not allowany kind of lateral fluid movements, the fluid flow rate is constant alongthe column. In real profiles, the goethite zone density is so low that itcollapses under the weight of the profile to a near constant value(Golightly, 1981).

The total computing time for 1-D transport calculations is defined as250,000 pore volumes of filling solution that are moved through thecolumn (107 shifts of solution per 40 column cells). The real geologicaltime in years could be estimated from residence time of solution in acell multiplied by number of solution shifts. Since the residence timeper cell was defined as 100,000 s, it will correspond to a summarycomputing time of about 10 Ma. We take this value as a connection ofour results to real geological time but it should be noted that in realitythe residence time will greatly change (decrease) due to porositychanges, formation of fractures and lateral dissipations of fluid.Therefore, this value of 10 Ma is an upper limit and certainly over-estimated. Further planned investigations will require implementationof this code into a 2D/3D fully coupled hydrological model to resolvethis drawback.

2.2. Modelling the geochemical system

Olivine composition in the model was selected according toTrescases (1975) as following: Mg1.82 Fe0.17 Ni0.01 Al0.006 SiO4. Olivinedissolution is assumed to be kinetically controlled whereas the pre-cipitation of secondary weathering products is considered to occur ac-cording to local equilibrium. Pokrovsky and Schott (2000) provided intheir work the following steady state specific dissolution rate for olivineat 25 °C in CO2-free solutions:

= − −r pHlog( ) 0.5 6.64 (2)

Which is consistent with an Arrhenius equation of the form of:

= −+r Aa E RTexp( / )Hn

a (3)

where r signifies the specific dissolution rate of olivine in (mol⋅m−2

⋅s−1), +aHn is the activity of protons (mol⋅l−1), A refers to a pre-ex-ponential factor, Ea designates an activation energy, R represents thegas constant, and T denotes absolute temperature.

Generally, the effect of atmospheric CO2 on olivine dissolution rateis very weak in acid to neutral solutions but may be important in al-kaline conditions. Results of Wogelius and Walther (1991) show a de-crease of the dissolution rate at elevated pH and low CO2 pressure whileexperiments conducted under similar conditions by Pokrovsky andSchott (2000) and Golubev et al. (2005) did not account such a decreasein comparison with their CO2-free experiments. Current modelling wasperformed according to the Eq. (2) and without taking into account theeffect of presence of CO2.

The overall olivine reaction rate in the model is as follows:

⎜ ⎟= ⎛⎝

⎞⎠⎛⎝

⎞⎠

R r AV

mmd

n0

0 (4)

where Rd signifies the overall reaction rate (mol⋅l−1 ⋅s−1), r is thespecific dissolution rate (mol⋅m−2 ⋅s−1) extracted from Eq. (2), A0

refers to the initial surface area of the solid (m2), V designates the vo-lume of solution (l), m0 represents the initial moles of solid, and m is themoles of solid at given time.

The factor (m/m0) n accounts for changes in surface area duringdissolution of olivine. For a monodisperse population of uniformlydissolving spheres or cubes, n = 2/3, since m is proportional to thevolume, or r3 (here r is the radius of the sphere or the side of the cube),while the surface area is proportional to r2.

The rate expression is embedded in the PHREEQC third orderRunge-Kutta-Fehlberg algorithm (Parkhurst and Appelo, 2013). Equi-librium phase composition is calculated before a kinetic calculation isinitiated and after a kinetic reaction increment. Calculations were doneat 25 °C with the code PHREEQC associated with the llnl.dat thermo-dynamic database (Johnson et al., 2000), that has been edited in orderto account for garnierite minerals. The reactions, involved in the pre-cipitation or dissolution of Mg and Ni end-member phyllosilicates aregiven in Table 1.

Basically, Mg-Ni-phyllosilicates are parts of solid-solution seriesextending from Mg to Ni end-members (e.g.Springer, 1974; 1976;Reddy et al., 2009) and effect of them being present as an ideal solidsolutions instead of separate minerals will be discussed. The process ofsubstitution of Ni for Mg in the lattice of serpentine is not consideredhere due to the lack of thermodynamic data needed to perform suchkind of modelling. Trescases (1975) demonstrated that many groundwaters from the base of the nickel laterite profiles in New Caledonia aresaturated or supersaturated with respect to talc-like mineral but rarelysaturated in serpentine. For the moment serpentine has been seldomreported to form under earth-surface conditions and, thus, suppressedfrom equilibrium phases that were allowed for precipitation. Instead, innature the precipitation rates favor the formation of metastable sepio-lite and kerolite in Al-poor environments (Stoessell, 1988). Thus, fortransport calculations at 25 °C the list of secondary weathering productsis following: Ferrihydrite, Gibbsite, Saponite-Mg, Kerolite, Pimelite,Sepiolite, Falcondoite, Quartz. Ferrihydrite is known to be an initialprecursor phase in the formation of goethite but is difficult to beidentified using standard characterization techniques (such as PXRD)due to its poorly crystalline nature. According to Cornell et al. (1992)the introduction of nickel to the system increases the stability of theferrihydrite phase, inhibiting its transformation to goethite and alsosuggesting that ferrihydrite could be still present in nickeliferous la-terites. Goethite and hematite form later on from ferrihydrite and theratio and crystallinity of these products are strongly influenced by pH.Schwertmann and Murad (1983) showed that at 24 °C maximum he-matite was formed in between pH 7 and 8, while maximum goethite atpH below 4 and above 12. This way, relying on a diagram of pH changewith depth, one can suppose what will be the final products in each partof profile.

The surface complexation model of Dzombak and Morel (1990),that takes into account binding of metals and protons on both strongand weak sites of ferrihydrite, which develops a charge depending onthe ions sorbed, has been used in order to simulate adsorption process.The adsorption constants required to parameterize the model for Niadsorption at the surface of ferrihydrite were obtained by fitting ex-perimental data (Dzombak and Morel, 1990) and are given in Table 2.

Two sites are defined for a diffuse-double-layer: i) Hfo_s, which is

Table 1Dissolution reactions for Mg and Ni end-members of garnierite and Saponite-Mg, with the corresponding equilibrium constants at 25 °C and 1 bar.

Mineral Reaction Log K

Kerolite Mg3Si4O10(OH)2:H2O+6H+=3Mg2++4SiO2+5H2O 25.79 (Stoessell, 1988)Pimelite Ni3Si4O10(OH)2:H2O+6H+=3Ni2++4SiO2+5H2O 11.46 (Nriagu, 1975)Sepiolite Mg4Si6O15(OH)2 : 6H2O+8H+=4Mg2++6SiO2+11H2O 30.44 (Stoessell, 1988)Falcondoite Ni4Si6O15(OH)2 : 6H2O+8H+=4Ni2++6SiO2+11H2O 12.31 (Nriagu, 1975)Saponite-Mg Mg3.16Al0.33Si3.67O10(OH)2+7.32H+ = 26.25 (llnl.dat)

3.16Mg2++0.33Al3++3.67SiO2+4.66H2O

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

277

Page 174: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

strong binding site, and ii) Hfo_w, which is weak binding site. Dzombakand Morel (1990) used 0.2 mol weak sites/mol ferrihydrite and0.005 mol (2.5% of weak sites) strong sites/mol ferrihydrite with asurface area 5.33×104 (m2 ⋅mol−1) and a weight of 89 g Hfo/mol Fe.To be consistent with their model, the relative number of strong andweak sites is kept constant as the total number of sites varies. In re-active 1-D transport modelling presented in this paper oxi-hydroxidesurfaces change in proportion as the ferrihydrite dissolves or pre-cipitates.

It is worth noting that the llnl.dat database contains thermodynamicdata for sorption of many trace metals. Once surface Ni complexes,Hfo_s and Hfo_w, are introduced in the model, the effect of other ca-tions, such as Mg2+, Fe2+, Ca2+ can be accounted for. Cations com-petition for sorption at the surface of ferrihydrite will, thus, influencethe mobility of leaching metals. In order to analyse this competition atransport modelling through one-cell with flux-open boundaries wasperformed. The input of one-cell model is identical to the main one with40 cells. This model has been used in order to reduce calculation timeand gain insight into the aforementioned competition of metals forsorption sites and precipitation of Ni-goethite solid solution with 2 wt%of Ni. It should be noted that in a results and discussion section wealways use the name goethite in spite of the fact simulations are donewith ferrihydrite constants (same chemical formula and different so-lubility constant value). This latter choice is applied to avoid any kindof confusions since ferrihydrite is being more geochemical rather thangeological term.

3. Results and discussion

3.1. 1-D reactive transport model

Fig. 4 shows the evolution of the paragenetic sequence and thedevelopment of the secondary nickel ores as a function of time. Theresults represent mass fractions of secondary weathering productsformed after 50,000 PV, 150,000 PV and 250,000 pore volumes (PV) offilling solution that are moved through the column and corresponds to2 Ma, 6 Ma and 10 Ma, respectively. The reactive transport model re-produces both the formation of the laterite profile in time with thick-ening of iron-rich zone and change of the Ni content with depth. Here,nickel supplied in the system by olivine dissolution was supposed to beredistributed in between solution, adsorption on goethite and dissolu-tion/precipitation of Ni-bearing minerals. In the following chapter ofdetailed understanding of Ni behavior during weathering is displayed.As can be seen from Fig. 4, the obtained chemical profile clearly dis-tinguishes the ferruginous/aluminous residual part where Si, Mg, andNi are leached, from the zone where silicates start to precipitate. Theeffect of silicate weathering on the water chemistry is primarily theaddition of cations (Ni2+, Mg2+, Fe2+, Fe3+; Al3+) and silica. pHvariation is balanced on one hand by silicate weathering reactionswhich consume H+ and on the other hand by leaching of the elementsout of the cell and renewing with low pH meteoric water which tend todecrease the pH value. Migration of the pH front, subdividing slightlyacidic at the surface to alkaline pH at depth, appears to have the maininfluence in controlling the precipitation/dissolution of neo-formed

silicate minerals, and, thereby, regulating the transport of elements inthe profile. The limonite zone is represented by goethite and minorprecipitations of gibbsite. Beneath, after the pH front, the secondaryweathering products are represented by falcondoite (Ni bearing phase),and massive precipitations of sepiolite and Saponite-Mg. Consideringthe results after 2 Ma, one can observe that part of olivine mass is notyet dissolved, providing pH up to 11 at the bottom of the column, andthat pimelite (Ni talc-like) precipitation (indicated by a narrow blackline on Fig. 4) occurs at depth, after the pH front. Pimelite precipitationis no more observed after 4 Ma. Indeed, after 4 Ma, olivine disappearsand the maximum pH value decreases from 11 to 9.5 after the pH front.

It worth noting that in New Caledonia the silicates, precipitated infractures or in a form of concentric zoning of Ni-bearing silicates(Cathelineau et al., 2016), are mostly represented by talc-like minerals.This is in contrast to sepiolite-like phases (falcondoite, sepiolite), ob-tained in current numerical modelling which is more common for Fal-condo Ni-laterite deposits observed in Dominican Republic (Villanova-de-Benavent et al., 2014). In order to find out why sepiolite-like pre-cipitated instead of talc-like minerals let us turn to the literature. Ac-cording to Stoessell (1988) the difference in solubility between keroliteand sepiolite is small enough that precipitation kinetics is probablydecisive in their formation. The metastability of kerolite relative tosepiolite, determined in the study of Stoessell (1988), was suggested byJones (1986). He reported kerolite as a precipitate at Stewart springs(California, USA) from waters with pH above 12, but with aqueous si-lica concentrations below (25 °C and 1 bar) quartz saturation. Jones(1986) suggested that with increasing pH and Mg content in water anddecreasing silica content, kerolite should precipitate at the expense ofsepiolite. Same conclusion might be reached for pimelite (Ni talc-like)and falcondoite (Ni sepiolite) precipitation since they are the Nicounterparts of kerolite and sepiolite, respectively.

It is consistent with the numerical results (Fig. 4) where, as it wasmentioned above, pimelite (narrow black line, 2 Ma column) pre-cipitated in the deeper part of profile where high pH of about 11 wasmaintained by dissolution of still presenting olivine. In this way it couldbe concluded that for precipitation of talc-like minerals at 25 °C in Al-poor environments it is necessary and sufficient to have an ampleamount of olivine in a system.

It should be noted that in real profiles olivine would persist forlonger time due to the different character of the porosity distribution. Infact, the extreme instability of olivine results in its replacement by si-licate phases in situ with the cations in the solutions diffusing outwardsto joints, where the water is flowing downwards. Because of this theolivine in the bulk material does not disappear before 6 million years asin the model. Nevertheless, an accurate representation of the processwill not be achieved until 2D modelling is attempted.

Returning to the results of the model the entire dissolution of olivinein the column after 6 Ma and 10 Ma (Fig. 4) led to a change in porewater pH with values of 5.6 close to the surface to about 9.5 at thebottom. Present day water analyses at Koniambo (northwest of NewCaledonia) indicated similar pH condition: from 5.5 near the surface toabout 10 at depth, either in the laterite or in the serpentinized peri-dotite (Jeanpert and Dewandel, 2013). Concerning the mass fraction ofNi bearing sepiolite (falcondoite), its accumulation and downwardmovement with pH front migration can be observed. The increase of Nicontent with time and depth is shown in Fig. 5.

It should be noted that the Ni concentration in wt% here is relatedto initial mass of olivine in a cell, serving as reference in order to beable comparing it through time. Normalizing to the current mass ofminerals presented in a cell adds a complexity in following comparisonof Ni enrichment since this mass is changing due to the leaching.Despite the fact that in the Fig. 5 at 250,000 PV (10 Ma) maximumconcentration of nickel (4.8 wt% ) seems to be lower than in previousstep (6 wt% after 200,000 PV, or 8 Ma), it appears that the thickness ofthe Ni rich zone is also higher after 250,000 PV than after lower PVcirculation. Therefore, the precipitation of Ni in the two closest cells

Table 2Thermodynamic data for a diffuse-double-layer surface “Hfo” (Hfo stands for hy-drous ferric oxide, i.e. ferrihydrite), derived from Dzombak and Morel (1990).

Reaction Log k

Hfo_ sOH + H+ = Hfo_ sOH+2 7.18

Hfo_ sOH = Hfo_ sOH− +H+ −8.82Hfo_ sOH + Ni2+ = Hfo_ sONi+ + H+ 0.37Hfo_ wOH + H+ = Hfo_wOH+

2 7.18

Hfo_ wOH = Hfo_ wOH− + H+ −8.82Hfo_ wOH + Ni2+ = Hfo_ wONi+ + H+ −2.5

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

278

Page 175: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

after 250,000 PV (10 Ma) (Fig. 4) with Ni content of 3.06 wt% and4.80 wt% at 9.25 m and 9.75 m of depth respectively, yields an overallequivalent Ni content of 7.89 wt%. This value is much higher com-paring to typical New Caledonian profile presented in Fig. 1 whichexhibits up to 2–3 wt% of Ni distributed throughout limonite and sa-prolite zone. The cause of it might be related to precipitation of silicatesat local equilibrium in our model, implying that simultaneous pre-cipitation occurs once favorable conditions are reached in a cell andresulting in a narrow zone of Ni precipitation. Also, in nature, nickel,released due to peridotite dissolution, may be partly lost by lateralmovement in solution.

The overall concentration of nickel of 7.89 wt% at the last step ofsimulation (10 Ma) represents 97.2% of the initial nickel content in thecolumn and distributed as follow: i) 4.27 wt% of Ni sorbed on ferri-hydrite surface, ii) 95.7 wt% precipitated with Ni-silicate (falcondoite)and iii) 1.69×10−4 % in solution along the whole column. From Fig. 5one can observe the movement of transition zone from the limonitehorizon to the saprolite, which is marked by a sharp increase in MgOcontent from 0 to up to 20 wt%. Iron content meanwhile varies fromvalues of about 65 wt% before the reaction front down to 5 wt% after it.While the iron has been oxidized, the nickel remains bivalent and issomewhat mobile, more so than iron, less so than magnesium. Thischemical distribution of elements is consistent with Fig. 1 and clearlydistinguishes the ferruginous residual part where Si, Mg, and Ni areleached and Mg discontinuity zone appears.

Nevertheless, in Fig. 1 Ni concentration is increasing throughoutlimonite zone up to 2 wt% at the contact with saprolite, while in nu-merical modelling this zone comes out to be free of Ni. According to themodelling results, adsorbed nickel appears to be governed by pH front.Being initially retarded compared to fluid movement, adsorbed Ni hasbeen eventually liberated into pore water at pH as low as 5.6 due toreaction front downward advancement upon silicates dissolution. De-pendence of nickel sorption on pH along with competitive retardation isdiscussed in the following chapter, devoted to a detailed understanding

of the weathering process. Yet, it could be concluded that adsorptioncan not explain such a high nickel content in limonite nowadays sincethe actual pH of soil waters can be in some instances as low as 4.5 dueto dissolved organic acids, thus, suggesting that nickel is held here bystronger ties. The latter might be explained by isomorphous substitutionof Ni for Fe in goethite and either significant process in soils related tospecific binding of elements to the variable charge surfaces of organicmatter. Adsorption on organic matter is similar to that on ferrihydritesince surface charge likewise depends on pH and solution composition,but the difference is that the sorption edge for organic matter is shiftedto lower pH. It implies that at low pH from about 4 to 7 adsorption onorganic matter will be dominating, but at higher pH the major role willbe passed to adsorption on oxyhydroxides. Present study does notconsider the issue of DOC due to its low concentration in well-drainedand scarce in terms of soil terrains of New Caledonia.

3.2. Detailed understanding of the weathering process

3.2.1. Effect of pimelite/kerolite being present as solid solutionThe numerical simulations presented above consider the Mg-Ni-

phyllosilicates to be pure end-members although in reality they arebeing members of solid-solution series extending from Mg and Ni end-members (e.g.Springer, 1974; 1976; Reddy et al., 2009; Villanova-de-Benavent et al., 2014; Cathelineau et al., 2015). Usually solid solutionsshow deviations from ideal behavior and the activity coefficients, thataccount for non-ideality, become a function of the excess free energy ofmixing. The free energy of mixing might be obtained from differentnon-ideal thermodynamic properties such as variations in fractionationfactor, miscibility gaps etc. (Glynn, 2000). Despite the fact that manyresearchers describe Mg-Ni-phyllosilicate members of solid solutions,the possibilities of modelling such a system remain poor due to the lackof thermodynamic data, thus, limiting simulations by assumption ofideality. In this way, for example, Galí et al. (2012) have calculatedLippmann diagrams for the garnierite solid solutions. This subchapter

Fig. 4. Mass fraction of secondary weathering products after 50,000 PV (2 Ma), 150,000 PV (6 Ma) and 250,000 PV (10 Ma). Yellow dotted line represents current pH front in the column.(For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

279

Page 176: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

aims at understanding the effect of e.g. pimelite and kerolite beingpresent as a solid solution instead of separate mineral phases. Linearcombination of reactions for kerolite and pimelite, presented in Table 1results in Eq. (5):

+ = ++ +Mg Si O (OH) : H O 3Ni Ni Si O (OH) : H O 3Mg3 4 10 2 22

3 4 10 2 22 (5)

At first, let us consider a situation where a solution is in equilibriumwith kerolite and pimelite as separate minerals. The mass action ex-pression for Eq. (5) looks as follows:

= × = =+

+K KK

[Ni Si O (OH) : H O][Mg Si O (OH) : H O]

[Mg ][Ni ]

10Kerolite

Pimelite

3 4 10 2 2

3 4 10 2 2

2 3

2 314.33

(6)

Then, in Eq. (6) the activities of two pure solid phases will be equalto one. Therefore, the ratio

+

+[Mg ][Ni ]

2 3

2 3 in solution is fixed to 1014.33 and

=+

+ 59, 795[Mg ][Ni ]

2

2 . The concentration of [Ni2+] is much smaller than of[Mg2+] because the solubility of pimelite is much smaller than ofkerolite.

In the second situation [Ni2+] is incorporated in the kerolitestructure as an ideal solid solution, which may be written asNiλMg3−λSi4O10(OH)2 : H2O, where λ represents the nickel molefraction. In this way, activities of kerolite and pimelite in Eq. (6) will nolonger be equal to one as for separate minerals, but will depend on λ.Therefore, if Mg concentration remains the same as before, then the Niconcentration must now be lower. It can clearly be seen in Fig. 6, thatshows the effect of Ni substitution in pimelite/kerolite solid solution onthe relative Ni concentration decrease in solution. Therefore, thecommon values of Ni mole fraction in New Caledonian pimelite/kero-lite solid solutions, i.e. around λ = 0.3–0.4, will lead to a decrease of

the Ni concentration by about 20% from the case with pure phases onassumption that Mg concentration remains the same (Fig. 6).

Thus, to conclude it should be said that the formation of solid so-lution lowers the aqueous concentration of the minor component, in ourcase nickel, but does not have a drastic effect on the global Ni massbalance. The latter means that despite the use of separate mineralphases our results of simulations may be considered as reliable.

3.2.2. Ni in solid solution with goethiteIt has already been mentioned above that sorption alone cannot

explain increase of Ni content throughout the limonite zone (Fig. 1). Inour 1-D model all Ni, adsorbed on goethite, was released in solution and

Fig. 5. Migration of pH front along the profile with corre-sponding Ni, Mg and Fe concentrations in the secondaryweathering products.

Fig. 6. Effect of Ni substitution in pimelite/kerolite solid solution (λ) on the relative Niconcentration decrease in solution (N).

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

280

Page 177: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

shifted at depth once the pH decrease below six after the front. Ob-viously, this latter phenomenon is not observed in the field, where Nistill persists in the limonite zone. According to Dublet et al. (2012),nickel content up to 2 wt% is common for New Caledonian laterites,where Ni2+ is associated with goethite via isomorphous substitution forFe3+. In order to account for this gap, here, we perform a sensitivitytest that considers additionally to our previous transport model (Fig. 4)a precipitation of Ni-bearing goethite. In this way, reaction for pre-cipitation of Ni-bearing goethite has been added in the thermodynamicdatabase as follows:

+ = + ++ + +Fe Ni OOH 3H 0.98Fe 0.03Ni 2H O0.98 0.033 2

2 (7)

The influence of such a small amount of Ni in structure of goethiteon its equilibrium constant has been neglected. Indeed, this constant,according to Eq. (7), will be defined as:

= × + × ++ +Klog 0.98 log [Fe ] 0.03 log [Ni ] 3pH3 2 (8)

Considering that the Ni concentrations in the model are as low as10−9 to 10−6 moles per liter, and taking into account Eq. (8) one cansee that the effect of Ni presence on the equilibrium constant will beminor.

Fig. 7 (A) shows the chemical distribution of the elements in mi-nerals versus pore volume (one million of PV represents one millionyears). Current reactive transport modelling has been performedthrough one cell with flux-open boundaries instead of 40 cells column.Incoming solution, olivine content and dissolution rate here are iden-tical to those used in 1-D transport model previously presented andsuch a test case was done in order to get detailed understanding of theweathering processes in each cell. X-axis in Fig. 7 represents pore vo-lumes, which is a number that shows how many times filling solutionmoved through the cell. Therefore, this axis represents time of evolu-tion of the bedrock until the formation of oxide-rich zone. This way theprincipal shape Ni enrichment curve in Fig. 7 (A) can be easily com-pared to Fig. 1. Indeed, we can highlight three zones for Ni enrichment:i) a zone of plateau where Ni co-precipitated with goethite, followed atpH higher than six by ii) a peak and iii) a decrease up to the value of Niconcentration in parent rock. Of course, since we investigate a cellstand alone, it has not any income of the elements from the upper cellsas it was in a column modelling, but the principal shape of Ni enrich-ment curve will remain the same with some differences in values. Thisshape, shown in Fig. 7 (A), is very close to that presented in Fig. 1.Similarly as it is observed in the field, the zone of residual laterites(Fig. 7 (A)) is mostly composed of iron oxy-hydroxides, and Ni, beingmore mobile, reaches its maximum concentrations at the base of thisregion. This area is marked by a sharp increase in pH and subsequentlyMgO content (Fig. 7 (A)). The latter means that pH appears to be themost important chemical parameter that drives the enrichment of

nickel in the weathering profile. Nevertheless, some differences can bespotted in between Fig. 7 (A) and Fig. 1 chemical distributions. Firstand foremost, as one can see, oxide Ni deposits show a regular decreaseof bulk nickel content towards the surface (Fig. 1) (e.g. Trescases, 1973;Wells et al., 2009; Dublet et al., 2012), while Fig. 7 (A) with themodelling results represents a stable plateau with value of 2 wt% of Ni.Dublet et al. (2012) have explained it by ageing of goethite in situthrough successive dissolution and precipitation cycles during later-itization. Due to a strong mismatch with Fe3+ both in terms of ionicradii and valence, nickel, substituted for Fe in goethite, is likely to beexpelled during each dissolution/precipitation cycle which leads topurification and increase of crystallinity of goethite. In addition, such adecrease in Ni content might be related to the fact that some part of Niin nature is also bonded in Mn-oxides (up to 10%). As far as these Mn-oxides are more soluble than iron oxy-hydroxides, they have beenleached from upper limonite zone and, therefore, a consequent drop inNi grade appears.

Another difference lies in a form of a peak and distribution of Niconcentrations around it. Here, there is a number of possible processesthat might have influenced the peak shape making it smoother andwider, namely: diffusion, dispersion, convective flows, kinetics of pre-cipitation of silicates, etc. It is important to note that the shape of nickelenrichment curve represents itself a superposition of different processesof Ni retention that are all evolving in time due to chemical conditionschange. In our simulations three different concurrent fates of Ni de-position in a profile were taken into account: i) Ni in a goethite crystallattice, ii) Ni sorbed on weak and strong goethite sorption sites, and iii)Ni precipitated with silicates (garnierite). Therefore, the superpositionthat represents the Ni enrichment curve (red line) in Fig. 7 (A) standsfor the sum of these three processes. The contribution of each process inthe total Ni enrichment is shown in Fig. 7 (B).

One can see that nickel mineralization begins with the formation ofNi-silicates at high pH, at the moment when ultramafic bedrock juststarts to dissolve. At the same time the precipitation of Ni-goethiteoccurs, with subsequent adsorption of Ni on its surface. Due to theleaching of the elements out of the cell in course of olivine dissolutionand refilling it by slightly acidic meteoric water at each time step, thepH in a modelled cell tends to decrease. Thus, according to the Fig. 7(B), at pH value above 9, Ni retention triggered by silicates precipita-tion prevails over the other processes. Then, the retention of nickel ingoethite lattice becomes increasingly important with the maximumachieved at pH 6 and lower. In between these two important mechan-isms, there is another one, related to adsorption of Ni on goethite. Thisprocess has its specific place during the weathering (Fig. 7 (B)); it ap-pears where pH is high enough to have sufficient charge of oxide sur-face to enhance sorption for binding of nickel. At the same time, at thevalues of pH of 8.5 or higher, it seems to be inhibited by some other

Fig. 7. Panel A represents the mobility of the elements in a 1-D reactive transport modelling through one cell; supergene enrichment of Ni and its retention by three different mechanisms:i) Sorption on goethite surface, ii) Retention in goethite structure, and iii) Precipitation with Ni-silicates. Panel B shows the split (redistribution) of Ni between different retentionmechanisms upon the formation of laterite. (For interpretation of the references to color in this figure, the reader is referred to the web version of this article.)

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

281

Page 178: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

process and subsequently replaced by precipitation of Ni-silicate (Fig. 7(B)). Therefore, adsorption is being rather a “transitional ore” sincebeing released from silicates the adsorbed nickel further passes to thelattice of goethite or precipitates with it. As it has been noticed pre-viously, the pH front in the nature most likely represents smoothercurve due to diffusion, dispersion, convective flows etc., and thewindow of existence of such a “transitional ore” on the surface of oxidesmight be geometrically larger. The following subchapter is fully de-voted to understanding of adsorption process of nickel alone in batchcalculations, as well as in competitive environment of transport mod-elling.

3.2.3. Competition of elements for sorption sitesAccording to the results of the 1-D numerical transport modelling,

presented in Chapter 3.1, adsorption of nickel strongly depends on pH(Figs. 5 and 7(B)) and does not appear in limonite zone due to fulldissolution of silicates and decrease of pH to the value of rainwater (i.e.5.6). Indeed, the surfaces of oxides carry a charge that enhances sorp-tion of metals and depends on pH and composition of the solution.Fig. 8 represents a batch calculation of nickel distribution among theaqueous phase and strong/weak sorption sites of 0.09 g of goethite,while Fig. 9 shows the shape of the sorption edge of Ni2+ as a functionof pH. These calculations are based on the Gouy-Chapman model from

Dzombak and Morel (1990) and show the results of distribution for low(10−8 M) and high (10−4 M) nickel concentrations. Concentrations ofNi, representative in our modelling, generally lie in this interval, beinglower than 10−6 M.

As one can see from both Figs. 8 and 9, nickel is more stronglysorbed at high pH values than at low pH values. The shape of thesorption edge is shown in Fig. 9 and actually depends on the Ni con-centration. At low concentrations (e.g. 10−8 M) the pH domain whereNi is sorbed to the surface of goethite increases from pH 8 to 10.5 whileat higher concentrations it reduces, general shape of the Ni sorptionedge smooths, and the acid branch shifts to the higher pH.

This is related to the fact that adsorption of high Ni concentrationsalso occurs on weak sites, what can be seen from the Fig. 8. Indeed, thedistribution of Ni among the solution and sorption sites of goethite doesalso depend on the Ni concentration. At low concentrations of nickel(Fig. 8 (A)) the strong binding sites outcompete the weak binding sitesover the entire pH range. Therefore, at high pH, most of the nickelresides at the strong binding sites. In case of larger nickel concentra-tions (Fig. 8 (B)), the strong binding sites prevail only at low pH.

Surface complexation theory appears to be a useful tool, helping toobtain a conceptual understanding of metal sorption behavior on mi-neral surfaces. While previous model (Figs. 8 and 9) was oversimplifiedrepresenting closed system, fixed goethite amount, and showing asorption just for Ni, assumed to be the only metal in solution, the fol-lowing one deals with competition of metals for sorption sites in a realcase. The phenomenon is highlighted in Fig. 10, where reactive trans-port modelling has been performed through one cell with flux-openboundaries as in previous Subchapter 3.2.2.

Fig. 8. Distribution of Ni among the aqueous phase and strong/weak surface sites of goethite (ferrihydrite) as a function of pH. Picture on the left (A) corresponds to low Ni concentration(10−8 M), while B represents high Ni concentration (10−4 M). Additional Y-axis represents charge balance of the solution. Solution was prepared with 0.1 M NaCl, which permits to useHCl or NaOH as reactant to keep pH increasing, and 90 mg pre-equilibrated ferrihydrite.

Fig. 9. The sorption edge of (10−8 M) and (10−4 M) Ni2+ at the surface of 0.09 g ofgoethite (ferrihydrite) as a function of pH in 0.1 M NaCl solution.

Fig. 10. Competition of Mg and Ni for sorption sites of goethite, based on the 1-Dtransport through 1 cell as in a Subchapter 3.2.2.

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

282

Page 179: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

Once the goethite has been formed in the system, adsorption startsto regulate the mobility of metals and retarding their movement. As canbe seen in Fig. 10, Mg at some point starts occupying weak sites ofgoethite, attaining maximum at pH of about 9.2 where, as it was shownin Fig. 9, nickel, being alone in the system, would have attained itsmaximum sorption.

With the subsequent decrease of pH magnesium releases again into porewater, giving way to Ni adsorption, which has a maximum at pH of 8.5. Itappears that at higher values of pH the Ni adsorption is inhibited by Mgbinding, which competes with Ni for the sorption sites. At lower ones nickelis released to pore water due to the decrease of surface charge. In this way atpH equal 6 all Ni has left goethite sorption sites, which explains why no Nisorption occur in the 1-D column (Figs. 5 and 7) before pH front (pH 5.6).

Thus, to conclude, adsorption of nickel on oxy-hydroxide surfaces(goethite) should not be regarded for Ni alone due to the big influenceof other cations releasing during the ultramafic bedrock alteration, inparticular Mg. This competition of cations for the sorption sites is tolarge extent controlled by pH. Adsorption of Ni becomes important in arestricted range of pH around 8, quickly releasing at lower than neutralpH values to pore water and being minor due to competition with Mg athigher ones.

4. Conclusions

The formation of Ni laterite profile from ultramafic parent rock, at25 °C, has been simulated by means of Phreeqc one-dimensional reac-tion-transport code. The main conclusions are listed below:

• The downward progression of the pH front controls thickening ofiron-rich zone, explains the mobility of the elements and governs theNi enrichment.

• Precipitation of talc-like silicates at the expense of sepiolite-like at25 °C is favored by high pH due to dissolution of still persistingolivine.

• The use of separate silicate phases instead of their solid solutions inthe simulations may be considered as reliable since it does not havea drastic effect on the global Ni mass balance.

• Modelling of Ni co-precipitation with goethite explains the presenceof Ni in upper limonite zone, where adsorption cannot retain Nianymore due to acidic pH. The decrease of bulk NiO content to-wards the surface might be explained by ageing of goethite in situthrough successive dissolution and precipitation cycles during la-teritization and also by dissolution of Mn oxides.

• Adsorption is being rather a “transitional ore” since after its releasefrom silicates the adsorbed nickel further passes to the lattice ofgoethite or precipitates with it. The process is fully controlled by pHfront and its shape.

• Nickel is more strongly adsorbed at high pH values than at low pHvalues up to some limit, when alone in the system.

• Adsorption of Ni becomes significant in a narrow range of pH due tothe competition of Mg and Ni for sorption sites of goethite surface.Thus, it is of importance to take into account the competition be-tween cations for sorption sites during the ultramafic bedrock al-teration

Presented in this paper reactive transport modelling of secondarynickel ore development is rather homogeneous (constant temperature,1D flow) and aims at understanding the main chemical and hydro-logical influence on it. Further successive dissolution and precipitationcycles of the minerals during lateritization are very closely tied to in-herited ores, tectonics, formation of preferential pathways, lateralmovement, change of relief, higher temperatures, that might explainore distribution heterogeneities observed on many Ni-laterite sites.Therefore, 2D/3D reactive transport simulations will be conducted toassess these aspects based on a coupling approach between geochem-istry, hydrodynamics and heat and mass transfer.

Acknowledgments

This work has been supported by the French National ResearchAgency through the national program “Investissements d’avenir” withthe reference ANR-10-LABX-21-01 / LABEX RESSOURCES21.

References

Barrow, N.J.N.J., Gerth, J.J., Brümmer, G.W.G.W., 1989. Reaction kinetics of the ad-sorption and desorption of nickel, zinc and cadmium by goethite. 2. Modeling theextent and rate of reaction. J. Soil Sci. 40, 437–450.

Brantley, S.L.S.L., Chen, Y.Y., 1995. Chemical weathering rates of pyroxenes and am-phiboles. Rev. Mineral. 31, 119–172.

Brümmer, G.W.G.W., Gerth, J.J., Tiller, K.G.K.G., 1988. Reaction kinetics of the ad-sorption and desorption of nickel, zinc and cadmium by goethite. I. Adsorption anddiffusion of metals. J. Soil Sci. 39, 37–51.

Butt, C.R.M.C.R.M., Cluzel, D.D., 2013. Nickel laterite ore deposits: weathered serpenti-nites. Elements 9, 123–128.

Carvalho-e-Silva, M.L.M.M.L.M., Ramos, A.Y.A.Y., Nogueira Tolentino, H.C.H.C.,Enzweiler, J.J., Netto, S.M.S.M., Martins Alves, M.C.M.C., 2003. Incorporation of Niinto natural goethite: an investigation by X-ray absorption spectroscopy. Am Mineral88, 876–882. http://dx.doi.org/10.2138/am-2003-5-617.

Cathelineau, M.M., Caumon, M.C.M.C., Massei, F.F., Brie, D.D., Harlaux, M.M., 2015.Raman spectra of Ni-Mg kerolite: effect of Ni-Mg substitution on O-H stretching vi-brations. J. Raman Spectrosc 46 (10), 933–940.

Cathelineau, M.M., Quesnel, B.B., Gautier, P.P., Boulvais, P.P., Couteau, C.C., Drouillet,M.M., 2016. Nickel dispersion and enrichment at the bottom of the regolith: for-mation of pimelite target-like ores in rock block joints (Koniambo Ni deposit, NewCaledonia). Mineral. Deposita 51 (2), 271–282.

Cornell, R.M.R.M., 1991. Simultaneous incorporation of Mn, Ni and Co in the goethite (α -FeOOH) structure. Clay Miner 26, 427–430.

Cornell, R.M.R.M., Giovanoli, R.R., Schneider, W.W., 1992. The effect of nickel on theconversion of amorphous iron(iii) hydroxide into more crystalline iron oxides in al-kaline media. J. Chem. Technol. Biotechnol. 53, 73–79.

Davis, J.A.J.A., Fuller, C.C.C.C., Cook, A.D.A.D., 1987. A model for trace metal sorptionprocesses at the calcite surface: adsorption of Cd and subsequent solid solution for-mation. Geochim. Cosmochim. Acta 51, 1477–1490.

Domènech, C.C., Galí, S.S., Villanova-de Benavent, C.C., Soler, J.M.J.M., Proenza,J.A.J.A., 2017. Reactive transport model of the formation of oxide-type ni-lateriteprofiles (Punta Gorda, Moa Bay, Cuba). Mineral. Deposita 1–18.

Dublet, G.G., Juillot, F.F., Morin, G.G., Fritsch, E.E., Fandeur, D.D., Brown, J.G.J.G.,2015. Goethite aging explains Ni depletion in upper units of ultramafic lateritic oresfrom New Caledonia. Geochim Cosmochim Ac 160, 1–15.

Dublet, G.G., Juillot, F.F., Morin, G.G., Fritsch, E.E., Fandeur, D.D., Ona-Nguema, G.G.,Brown, J.G.J.G., 2012. Ni speciation in a New Caledonian lateritic regolith: aquantitative X-ray absorption spectroscopy investigation. Geochim Cosmochim Ac95, 119–133.

Dzombak, D.A.D.A., Morel, F.M.M.F.M.M., 1990. Surface Complexation Modeling:Hydrous Ferric Oxide. New York, John Wiley and Sons.

Fischer, L.L., Brümmer, G.W.G.W., J, B.N., 2007. Observations and modelling of the re-actions of 10 metals with goethite: adsorption and diffusion processes. Eur J Soil Sci58, 1304–1315. http://dx.doi.org/10.1111/j.1365- 2389.2007.00924.x.

Fletcher, R.C.R.C., Buss, H.L.H.L., Brantley, S.L.S.L., 2006. A spheroidal weatheringmodel coupling porewater chemistry to soil thicknesses during steady-state denu-dation. Earth Planet Sci Lett 244, 444–457.

Freyssinet, P.P., Butt, C.R.M.C.R.M., Morris, R.C.R.C., Piantone, P.P., 2005. Ore-formingprocesses related to lateritic weathering. Econ. Geol. 681–722 100th AnniversaryVolume.

Galí, S.S., Soler, J.M.J.M., Proenza, J.A.J.A., Lewis, J.F.J.F., Cama, J.J., Tauler, E.E.,2012. Ni-enrichment and stability of Al-free garnierite solid solutions: a thermo-dynamic approach. Clays Clay Miner 60, 121–135.

Gerth, J.J., Brümmer, G.W.G.W., Tiller, K.G.K.G., 1993. Retention of Ni, Zn and Cd by Si-associated goethite. Z. Pflanzenern. Bodenk. 156, 123–129.

Gleeson, S.A.S.A., Butt, C.C., Elias, M.M., 2003. Nickel laterites: a review. SEG Newsletter54, 9–16.

Glynn, P.D.P.D., 2000. Solid-solution solubilities and thermodynamics: sulfates, carbo-nates and halides. Rev.Mineral. 40, 481–511.

Golightly, J.P.J.P., 1981. Nickeliferous laterite deposits. Econ. Geol. 710–735 75th an-niversary volume.

Golightly, J.P.J.P., 2010. Progress in understanding the evolution of nickel laterites.Society of Economic Geologists Special Publication 15, 451–485.

Golubev, S.V.S.V., Pokrovsky, O.S.O.S., Schott, J.J., 2005. Experimental determination ofthe effect of dissolved CO2 on the dissolution kinetics of Mg and Ca silicates at 25 °C.Chem. Geol 217, 227–238.

Guilbert, J.M.J.M., Park, C.F.C.F., 1986. The Geology of Ore Deposits, 4th ed. New York :W.H. Freeman.

Herzberg, C.C., Asimow, P.D.P.D., Ionov, D.A.D.A., Vidito, C.C., Jackson, M.G.M.G.,Geist, D.D., 2013. Nickel and helium evidence for melt above the core-mantleboundary. Nature 493, 393–397.

Jeanpert, J.J., Dewandel, B.B., 2013. Analyse préliminaire des donées hydrogéologiquesdu massif du Koniambo. Public Report BRGM/RP-61765-FR. pp. 102.

Johnson, J.J., Anderson, G.G., Parkhurst, D.D., 2000. Database From thermo.-com.v8.r6.230 Prepared at Lawrence Livermore National Laboratory. (Revision:

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

283

Page 180: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite

1.11).Join, J.-L.J.-L., Robineau, B.B., Ambrosi, J.-P.J.-P., Costis, C.C., Colin, F.F., 2005. Système

hydrogéologique d’un massif minier ultrabasique de nouvelle-calédonie. ComptesRendus Geoscience 337 (16), 1500–1508.

Jones, B.F.B.F., 1986. Clay mineral diagenesis in lacustrine sediments. Studies in diag-enesis. U.S. Geol. Surv. Bull. 1578, 291–300.

Lebedeva, M.I.M.I., Fletcher, R.C.R.C., Balashov, V.N.V.N., L, B.S., 2007. A reactive dif-fusion model describing transformation of bedrock to saprolite. Chem Geol 244,624–645.

Lelong, F.F., Tardy, Y.Y., Grandin, G.G., Trescases, J.J.J.J., Boulange, B.B., Wolf,K.H.K.H., 1976. Pedogenesis, chemical weathering and processes of formation ofsome supergene ore deposits. Supergene and Surficial Ore Deposits. Texture andFabrics. Handbook of Strata-Bound and Stratiform Ore Deposits: 3. Elsevier,Amsterdam, pp. 93–133 v.

McKenzie, R.R., 1980. The adsorption of lead and other heavy metals on oxides of Mn andFe. Aust. J. Soil Res. 18, 61–73.

Navarre-Sitchler, A.A., Steefel, C.I.C.I., Sak, P.B.P.B., Brantley, S.L.S.L., 2011. A reactive-transport model for weathering rind formation on basalt. Geochim Cosmochim Ac 75,7644–7667. http://dx.doi.org/10.1016/j.gca.2011.09.033.

Nriagu, J.O.J.O., 1975. Thermochemical approximation for clay minerals. Am. Mineral.60, 834–839.

Parkhurst, D.L.D.L., Appelo, C.A.J.C.A.J., 2013. Description of Input and Examples forPHREEQC Version 3-A Computer Program for Speciation, Batch-Reaction, One-Dimensional Transport, and Inverse Geochemical Calculations. U.S. GeologicalSurvey Techniques and Methods, book 6, chap. A43. pp. 497 p., available only athttp://pubs.usgs.gov/tm/06/a43/.

Pokrovsky, O.S.O.S., Schott, J.J., 2000. Kinetics and mechanism of forsterite dissolutionat 25 °C and pH from 1 to 12. Geochim. Cosmochim. Acta 64, 3313–3325.

Quesnel, B.B., Boulvais, P.P., Gautier, P.P., Cathelineau, M.M., Cdric, M.J.M.J., Dierick,M.M., Agrinier, P.P., Drouillet, M.M., 2015. Formation of Silica and Magnesite Veinsin the Massif of Peridotite of Koniambo: Geometric and Stable Isotopes Data. 13thSGA Biennial Meeting 2015. Proceedings. 3. pp. 1189–1192.

Reddy, B.J.B.J., Frost, R.L.R.L., Dickfos, M.J.M.J., 2009. Characterisation of Ni silicate-bearing minerals by UV- vis-NIR spectroscopy. Effect of Ni substitution in hydrous Ni-Mg silicates. Spectrochim. Acta A Mol. Biomol. Spectrosc. 71 (5), 1762–1768.

Schultz, M.F.M.F., Benjamin, M.M.M.M., Ferguson, J.F.J.F., 1987. Adsorption and deso-rption of metals on ferrihydrite: reversibility of reaction and sorption properties ofthe regenerated solid. Env. Sci. Technol 663–669.

Schwertmann, U.U., Murad, E.E., 1983. Effect of pH on the formation of goethite andhematite from ferrihydrite. Clays and Clay Minerals 31 (4), 277–284.

Singh, B.B., Sherman, D.M.D.M., Gilkes, R.J.R.J., 2002. Incorporation of Cr, Mn and Niinto goethite (α -FeOOH): mechanism from extended X-ray absorption fine structurespectroscopy. Clay Miner. 37, 639–649.

Soler, J.M.J.M., Lasaga, A.C.A.C., 1996. A mass transfer model of bauxite formation.Geochim Cosmochimi Ac 60, 4913–4931.

Soler, J.M.J.M., Lasaga, A.C.A.C., 1998. An advection-dispersion-reaction model of

bauxite formation. J Hydrol 209, 311–330.Springer, G.G., 1974. Compositional and structural variations in garnierites. Can. Mineral.

12, 381–388.Springer, G.G., 1976. Falcondoite, nickel analogue of sepiolite. Can. Mineral. 14,

407–409.Stipp, S.L.S.S.L.S., Hochella, M.F.M.F., Parks, G.A.G.A., Leckie, J.O.J.O., 1992. Cd uptake

by calcite, solid-state diffusion, and the formation of solid-solution: interface pro-cesses observed with near-surface sensitive techniques (XPS, LEED and AES).Geochim. Cosmochim. Acta 56, 1941–1954.

Stoessell, R.K.R.K., 1988. 25 °C and 1 atm dissolution experiments of sepiolite and ker-olite. Geochim. Cosmochim. Acta 52, 365–374.

Swamy, Y.Y., Kar, B.B.B.B., Mohanty, J.K.J.K., 2003. Physico-chemical characterizationand sulphatization roasting of low-grade nickeliferous laterites. Hydrometallurgy 69,89–98.

Thom, J.G.M.J.G.M., Dipple, G.M.G.M., Power, I.M.I.M., Harrison, A.L.A.L., 2013.Chrysotile dissolution rates: implications for carbon sequestration. Appl. Geochem.35, 244–254.

Thorne, R.L.R.L., Roberts, S.S., Herrington, R.R., 2012. Climate change and the formationof nickel laterite deposits. Geology 40, 331–334.

Trescases, J.J.J.J., 1973. Weathering and geochemical behaviour of the elements of ul-tramafic rocks in New Caledonia. Bureau Mineral Res Geol Geophys Dep MineralEnergy Canberra Bull 141, 149–161.

Trescases, J.J.J.J., 1975. L’élvolution gélochimique supergèlne des roches ultrabasiquesen zone tropicale: Formation des gisements nickélifères de Nouvelle-Calédonie: Ph.D.thesis Mémoires ORSTOM (Office de la Recherche Scientifique et Technique Outre-Mer).

Troly, G.G., Esterle, M.M., Pelletier, B.B., Reibel, W.W., 1979. Nickel Deposits in NewCaledonia: Some Factors Influencing Their Formation. International LateriteSymposium, New Orleans,. pp. 85–119.

Ulrich, M.M., 2010. Péridotites et serpentinites du complexe ophiolitique de la Nouvelle-Calédonie. Université de la Nouvelle-Calédonie et Université de Grenoble Ph.D.thesis.

Ulrich, M.M., Muñoz, M.M., Guillot, S.S., Chauvel, C.C., Cluzel, D.D., Picard, C.C., 2011.Weathering Effects on the Mineralogical and Geochemical Composition of the NewCaledonia Ophiolite. Goldschmidt Conference.

Villanova-de-Benavent, C.C., Proenza, J.A.J.A., Galí, S.S., Garca-Casco, A.A., Tauler, E.E.,Lewis, J.F.J.F., Longo, F.F., 2014. Garnierites and garnierites: textures, mineralogyand geochemistry of garnierites in the Falcondo Ni-laterite deposit, DominicanRepublic. Ore Geol. Rev. 58, 91–109.

Wells, M.A.M.A., Ramanaidou, E.R.E.R., Verrall, M.M., Tessarolo, C.C., 2009. Mineralogyand crystal chemistry of garnierites in the Goro lateritic nickel deposit, NewCaledonia. Eur. J. Mineral. 21, 467–483.

Wilson, M.J.M.J., 2004. Weathering of the primary rock-forming minerals: processes,products and rates. Clay Minerals 39, 233–266.

Wogelius, R.A.R.A., Walther, J.V.J.V., 1991. Olivine dissolution at 25 °C - effects of pH,CO2, and organic-acids. Geochim. Cosmochim. Acta 55, 943–954.

A. Myagkiy et al. Chemical Geology 466 (2017) 274–284

284

Page 181: Genèsedesconcentraons de&NickelenNouvelle&Calédonie ......2001]. The Poya Terrane occurs as a package of steep fault-bounded sheets intercalated between the overlying Peridotite